You are on page 1of 13

Applied Catalysis B: Environmental 46 (2003) 293–305

Iron-containing silicalites for phenol catalytic wet peroxidation


Nicolas Crowther, Faı̈çal Larachi∗
Department of Chemical Engineering & CERPIC, LAVAL University, Ste-Foy, Que., Canada G1K 7P4
Received 8 December 2002; received in revised form 18 April 2003; accepted 21 May 2003

Abstract
Heterogeneous catalytic wet peroxidation (CWPO), involving the mineralization (i.e. complete conversion to CO2 and
H2 O) of organic compounds is a possible path for the treatment of toxic and/or bio-refractory wastewater streams. The aim of
the present work was to synthesize and characterize five solid-phase catalyst formulations featuring iron(III) species supported
on silica-based mesoporous molecular sieves of the MCM-41 and HMS-types. These materials were tested for the CWPO of
aqueous phenol in a batch slurry reactor using hydrogen peroxide, as an oxidant, under very mild conditions (P = 1 atm and
T = 80 ◦ C). The five materials were produced according to four recipes. Structural data was provided by XRD and nitrogen
adsorption–desorption, the iron content was assessed by atomic absorption. The freshly obtained catalysts, with embedded
templates within pores, were activated either by calcination or through solvent extraction, then used for CWPO. The spent
catalysts were recovered and characterized before thermal re-activation and re-exposure to fresh phenol-containing solutions.
For all materials, in the first run, the conversion of phenol was total and rapid (less than 15 min), and the removal of total
pollution plateaued between 55 and 85% in 180 min with mineralization selectivities of ca. 95%. The materials underwent
leaching of the iron, from 6% (w/w) to total. It was possible to perform successful CWPO reactions with some recycled
materials. In terms of activity as well as robustness in the reaction conditions, one catalyst was able to remove 81% of the
organic load with a selectivity of 93%, and simultaneously leaching 6% iron off. During the second run, it showed less activity
but endured 2% leaching only. Surprisingly, the structure of this best catalyst was not mesoporous nor microporous and
exhibited a very low specific surface area.
© 2003 Elsevier B.V. All rights reserved.
Keywords: Heterogeneous catalytic wet peroxidation; Wastewater treatment; Mesoporous silicalite; Iron; Hydrogen peroxide; Phenol; Leaching

1. Introduction carbon dioxide and/or other harmless species. Cata-


lysts are certainly expected to occupy a central place
The increasingly stringent water quality regula- in such processes [1–3]. Therefore, the design of new
tions and demand for water reuse after treatment catalyst formulations is actively pursued in industry
have stimulated strong research efforts aimed at the and academia to meet market demands driven by
development of cheap and efficient routes to destroy increasingly tight profit margins, especially because
toxic and/or bio-refractory organic compounds from no immediate dollar profits are expected in cleaning
liquid streams. One possible path is the complete “end-of-pipe” wastewater streams.
oxidation (or mineralization) of these compounds to Several successful processes can be applied at
various stages of water treatment, and undoubtedly
∗ Corresponding author. Tel.: +1-418-656-3566; catalytic wet oxidation (CWO), using pure oxygen or
fax: +1-418-656-5993. air for oxidant, has proven its worth for wastewater
E-mail address: faical.larachi@gch.ulaval.ca (F. Larachi). treatment, particularly when organic contaminants at

0926-3373/$ – see front matter © 2003 Elsevier B.V. All rights reserved.
doi:10.1016/S0926-3373(03)00224-8
294 N. Crowther, F. Larachi / Applied Catalysis B: Environmental 46 (2003) 293–305

low or moderate loads are concerned. CWO is usually the pH of the gel greatly influences the thickness of the
practiced between 100 and 300 ◦ C and at pressures pore walls. Gels with a high pH frequently possess thin
in the 1–10 MPa range. These severe conditions are and fragile walls, as opposed to thick ones obtained
required for minimizing the impact of CWO intrinsic for neutral pH.
main limitations which are the oxygen low solubil- There are two ways in which the iron can mod-
ity in water and the hindered oxygen supply to the ify the silica structure. The metal may simply be in
liquid-phase due to the gas- and liquid-side mass contact with the surface of the pore walls, and upon
transfer resistances. Circumventing completely the calcination, nanoclusters of octagonal crystalline iron
limited-solubility as well as the interfacial resistance oxide (Fe2 O3 ) are formed. This extra-framework iron
barriers of the oxidant implies increasing temperature is retained to the support only by weak Van der Waals
and pressure above the water critical point (374 ◦ C, bonds. The iron oxide clusters are prone to easily leach
218 atm) yielding limitless O2 -water miscibility and off and/or be washed out during the CWPO treating.
resulting in a collapse of the gas–liquid interface [4,5]. In the second way, the iron may well be incorporated
Another far more economical alternative route to su- by substitution into the silica framework, and thus be
percritical wet oxidation is to opt for a liquid-phase attached by covalent bonds. Iron lies in a tetragonal
oxidant, such as hydrogen peroxide, that would allow geometry and is referred to as framework iron. Iron can
the (wet peroxidation) conditions to decrease dramat- be incorporated in the silica framework through direct
ically to atmospheric pressure and sub-boiling water synthesis by introducing an appropriate iron precursor
temperature [6,7]. in the synthesis gel as done for other atoms such as
The use of wet peroxidation homogeneous catalysts titanium or boron [12]. In this way, iron atoms present
like the Fenton catalysts Fe2+ /Fe3+ , or other dissolved at the edge of the pore walls are likely to represent
transition metal cations, e.g. Cu2+ , Mn2+ , and Co2+ , resistant active centers.
has proven to be quite efficient [8,9], but this neces- Substitution is more difficult than simple contact
sitates a tight pH control to prevent precipitation and to the surface and there is often a mixture of both
extra steps for the recuperation and the reuse of the framework and extra-framework metal species when
catalyst [10]. The alternative is the development of substitution is intended. Some works report that iron
solid-phase catalysts, which feature efficiency as well oxide clusters color the materials in shades of brown,
as stability under the reaction conditions. whereas framework iron is colorless [13–15]. Finally,
The aim of the present work is to synthesize and a maximum limit of 0.9–1.0 wt.% for framework po-
characterize a series of heterogeneous catalysts for sitions was established by Wang et al. [15].
the catalytic wet peroxidation (CWPO) of organic Some works report that the presence in the synthe-
compounds in water. Phenol was chosen as a model sis gel of fluoride anions maximizes the proportion of
reactant for comparison with available literature on framework-substituted metals [16–18]. Besides being
CWPO. Hydrogen peroxide was chosen as the oxi- widely used as condensation agents in sol–gel synthe-
dant, having both advantages of being relatively in- sized silica [18], fluorides are thought to form soluble
expensive and a liquid-phase compound. The catalyst complexes with iron(III) at the pH values prevailing
features iron(III) species supported on silica-based in the synthesis [19].
mesoporous molecular sieves (MCM and HMS hexag- Previous studies in our group have shown that
onal structures). The expected characteristics are CWO catalysts may undergo some deactivation via
higher specific surface area together with large pore the obstruction of the catalytic sites by adsorption
size which is expected to enhance intraparticle dif- and/or polymerization of the organics present [20–22].
fusion when compared to microporous materials like The fraction of carbon actually converted to CO2 ,
zeolites. Iron doping was chosen for its cheapness as will be measured to evaluate the mineralization selec-
well as for its known participation in the conventional tivity of the synthesized materials. Deactivation can
Fenton/Haber–Weiss homogeneous reaction pathway. also occur via the leaching of the active species out
It has been shown that the final stability, especially of the material. This phenomenon may be attributed
the hydrothermal stability of the mesoporous network to several factors. The first of which is that the pH
depends on the synthesis conditions [11]. In particular of the CWPO reaction medium is acidic. Iron and
N. Crowther, F. Larachi / Applied Catalysis B: Environmental 46 (2003) 293–305 295

particularly extra-framework iron oxide are readily then a small amount of ethanol, and dried overnight
dissolved in an acidic environment. Secondly, strong at 80 ◦ C. The “fresh” form of the obtained catalyst A
complexing and solvolytic properties of the oxidant is denoted A-f.
and/or intermediate products (e.g. catechol) may rup- Catalyst B was produced in a similar way, except
ture the metal-oxygen bonds [23,24]. Finally, the pH that the iron potassium oxalate complex was replaced
range corresponds to the stability zone of Fe3+ in by an iron potassium EDTA complex, prepared from
aqueous solution [19]. an appropriate amount of K2 EDTA (EDTA 1:2 KOH)
combined with Fe(NO3 )3 ·9H2 O (Aldrich) in a 1:1 mo-
lar proportion in water.
2. Experimental Catalyst C was synthesized as reported by Pasqua
et al. [17]. A first solution was prepared from
2.1. Catalyst preparation 7.65 g of CTAB, 14.81 g of NH4 F and 197.1 g of
water. After solubilization, 2.02 g Fe(NO3 )3 ·9H2 O
The formation of hexagonal mesoporous silica with 65.7 g water were added. Then, 20.85 g of
structures, based on a sol–gel process, is a relatively tetraethylorthosilicate (TEOS, Aldrich) was added
complex phenomenon and has been thoroughly de- slowly. The gel had the following molar composition:
tailed in numerous articles since their discovery [25]. 0.1SiO2 ·0.005Fe·0.021CTAB·0.4NH4 F·146H2 O. Pre-
It implies a hydrothermal sol–gel process involving cipitation was readily observed, and the gel was left
the condensation of silica species (soluble or col- to stir for 4 h at room temperature, before being trans-
loidal), with uptake of other inorganic compounds if ferred to a Teflon® lined stainless steel autoclave to
present, on the surface of ordered micelles of surfac- age statically at 105 ◦ C for 24 h. The pH remained
tant acting as template. This eventually leads to the around 7.7 throughout the process. The precipitate
precipitation of the solid-phase which is recovered. was then filtered, washed with water and a dash of
Five iron-loaded hexagonal or lamellar mesoporous ethanol and left to dry at 80 ◦ C overnight.
catalysts were prepared with iron loading between 1.2 Catalyst D was prepared according to Reddy
and 16.6% (w/w). Samples A–D were expected to dis- and Song [26]. First, 4 g of Cab-O-Sil were dis-
play a MCM structure (ionic surfactant), and for sam- persed in 25.7 g of water. Secondly, 11.6 g of
ple E, HMS composition (neutral template). tetramethylammonium-silicate solution (Aldrich) was
Synthesis of catalyst A was performed using a mixed with 5.5 g of sodium silicate solution (Aldrich)
procedure described by Alves and Pastore [16]. then added to the Cab-O-Sil and left 15 min to stir.
First, a “silica source” was prepared by adjusting a Thirdly, a solution of 0.67 g of iron nitrate (Aldrich)
commercial sodium silicate solution (Aldrich 27% and 10 g of water was slowly added to the previ-
SiO2 , 14% NaOH, d = 1.390) to give a 90 g/l SiO2 ; ous mixture. Finally, a solution of 12.9 g of CTAB
180 g/l NaOH solution. To 66 ml of the silica source in 77 g of water was added to form the gel. The
brought to a temperature of 75 ◦ C approximately, was gel molar composition was: 1SiO2 ·0.025Fe·0.3124-
slowly added 12.7 g of a 33 wt.% aqueous solution of CTAB·0.1728NaOH·0.0876TMA·63.3H2 O. The gel
cetyltrimethylammonium bromide (CTAB, Aldrich) was allowed to stir for 2 h, and was transferred to the
under stirring. After further stirring for 30 min, the Teflon® lined stainless steel autoclave to age stati-
pH was slowly lowered to 12 with concentrated cally at 100 ◦ C for 24 h. The precipitate was filtered,
hydrofluoric acid (EM Science) used as received. washed and left to dry at 100 ◦ C overnight.
Then, approximately 72 ml of a 0.07 mol/l solution of Catalyst E preparation was similar to that detailed
iron(III) potassium oxalate (Alfa Aesar) was added in Ruel and Gontier [27]. A solution of 17.69 g TEOS,
to the gel. The molar composition of the gel mixture 24.76 g ethanol and 4.95 g isopropyl alcohol was
was 0.1SiO2 ·0.005Fe·0.011CTAB·100H2 O. After 4 h placed in a beaker. Slowly added to it was a second
of stirring, the final pH was around 11.6. The gel was solution of 1.5 ml aqueous HCl 1 M, 54.25 g water,
allowed to age in Teflon® lined stainless steel auto- 4.13 g dodecylamine (Aldrich) and 0.84 g iron nitrate.
claves at 150 ◦ C for 66 h. After this, the precipitate The gel had the following composition: 1SiO2 ·0.025-
was filtered, washed extensively with deionised water Fe·0.26DDA·6.3EtOH·0.97IPA·0.02HCl·36H2 O. Af-
296 N. Crowther, F. Larachi / Applied Catalysis B: Environmental 46 (2003) 293–305

ter stirring for 1 h, it was left to stand overnight, then The materials will be referred to by their letter
the precipitate was filtered, washed and left to dry at followed by -f, -c, -e, -r, respectively, for “fresh”,
100 ◦ C overnight. “calcined”, “extracted”, and “reused”.
All the materials described above are referred as
“fresh”. In this form, the template remained trapped 2.3. Oxidation reaction
inside the pores. In order for the catalyst to be ready for
use, the template was removed either by calcination, Phenol oxidation was carried out in a double-jacket
or by extraction with appropriate template solvents. glass batch reactor. The entire reactor volume was
For calcination, the samples were placed in an oven first sparged by flushing with industrial grade he-
and the temperature was raised from room to 550 ◦ C lium (Praxair), then closed and kept under slight He
in 6 h under nitrogen, and kept up for six more hours overpressure. An aqueous phenol solution at 0.260 g/l
under dry air. Extraction was performed by dispersing corresponding to 200 mg/l of carbon (ppm C) was
the fresh catalyst in a mixture of ethanol and water prepared by dissolving phenol (EM Science) into
under stirring, aided with small amounts of NaCl or deionized water brought to pH 3.5 with sulfuric acid
HCl, for 3–4 h. The extracted catalysts were recovered (EM Science), then sonically degassed. The deionized
by filtration, washing and drying at 80 ◦ C overnight. water was obtained from a Culligan Model E1 Plus
These materials then are referred to as “activated”. purification system (electrical conductance < 1 ␮S).
The total volume of liquid (phenol and hydrogen per-
2.2. Catalyst characterization oxide solutions) being 300 ml, the adequate amount
of phenol solution was poured into the reactor. The
To identify the structure and the distance between catalyst (0.15 g, i.e. load of 0.5 g/l) was then contacted
(1 0 0) diffraction lattice planes, X-ray powder diffrac- with the solution, and both were allowed to reach the
tion (XRD) patterns were recorded on a Siemens experimental temperature of 80 ◦ C in approximately
D5000 diffractometer using the Cu K␣ radiation at 30 min. The volume of 1 mol/l hydrogen peroxide
40 kV to 30 mA (λ = 1.54184 Å), at 1 ◦ /min [2θ] over commercial solution (EM Science) to add was cal-
the range 1–8◦ . The specific BET surface areas and the culated by considering the stoichiometric amount of
pore size distribution were determined using N2 ad- H2 O2 required to convert phenol to CO2 (14 mol per
sorption and the BET model. The nitrogen adsorption mol of phenol) plus an excess of 50–60% above the
isotherms at 77 K were measured on an OMNISORP stoichiometric needs. The reaction time reference was
100 or a QUANTACHROME Autosorp-1 device, af- taken when the volume of hydrogen peroxide solution
ter pre-treatment under vacuum at 250 ◦ C to a residual was injected in one shot into the reactor.
pressure of about 10−5 Torr. The total iron content Samples of the reaction medium were analyzed
of the calcined samples was measured by atomic in order to obtain the total organic carbon (TOC),
absorption spectroscopy (AAS), on a Perkin-Elmer the phenol and the hydrogen peroxide conversions,
AAnalyst100 or with a Perkin-Elmer Optima3000 and to monitor the pH evolution and the potentially
inductively coupled plasma spectrometer, ICP. The leached off iron from the catalyst. TOC of acidified
samples were first dissolved in an aqueous HCl–HF samples (to remove dissolved CO2 ) was measured
mixture. Catalysts were investigated for their carbon using a combustion/non-dispersive infrared gas ana-
content with a Carlo-Erba 1106 CHN analyzer and lyzer (model Shimadzu 5050 TOC analyzer, Mandel
a FT-IR spectrometer (Nicolet), allowing to evaluate Scientific Inc., Montréal, PQ).
the fraction of carbon removed from the liquid-phase Disappearance of phenol, the parent pollutant, was
but not converted to CO2 during CWPO. monitored using a high-performance liquid chromato-
Catalysts for reuse were recovered by filtration from graph (HPLC), which consisted of a combination
the reaction medium, then dried at 100 ◦ C overnight of model 510 Waters dual-solvent delivery system,
and underwent complete calcination cycle just like for a model 490 programmable multi-wavelength ultra-
the activation. The materials ready for reuse were an- violet absorbance detector, an automated gradient
alyzed for their structure by XRD, and for their total controller, and a Phenomenex Bondclone® C18,
iron content by AAS or ICP. 300 mm × 3.9 mm × 10 ␮m. The unit was operated
N. Crowther, F. Larachi / Applied Catalysis B: Environmental 46 (2003) 293–305 297

isocratically at 0.8 ml min−1 using as a mobile phase of one or the other of the reactants, in the experimen-
50:50 (volume ratio) water:methanol. UV detection tal conditions described previously. In the absence of
was performed at 254 nm. Peak area recording and iron, no significant consumption of phenol or hydro-
integration was achieved using a Hewlett-Packard in- gen peroxide was noticed. Homogeneous iron at a con-
tegrator (HP 3390 A). HPLC calibration was achieved centration of 12 ppm (equivalent to 0.5 g/l of catalyst
by using phenol concentrated solutions from which with 2.5% (w/w) iron loading) was contacted with a
daughter solutions were obtained by successive di- phenol-free solution of hydrogen peroxide. A 35% de-
lutions. The main intermediate oxidation products, crease in the amount of oxidant was observed in the
detected but not quantitated, were catechol, hydro- space of 1 h. Thus, free iron(III) cations were effective
quinone, quinones, muconic acid, maleic acid, oxalic in activating hydrogen peroxide.
acid and acetic acid [6]. H2 O2 consumption was Fig. 1 presents a typical time course of the relative
given by the iodine–thiosulfate titration method. The residual concentrations for total organic carbon, phe-
content of metal ions was determined by AAS or ICP, nol and hydrogen peroxide obtained for catalyst B-c.
described previously. The trends exhibited in Fig. 1 are also representative
of the catalytic behavior for the other catalysts A-c
through D-c.
3. Results Table 1 summarizes the TOC, phenol and peroxide
relative residues for all the catalysts, as well as figures
3.1. Oxidation reaction characterizing the performance of the oxidant. All cat-
alysts were able to completely oxidize the parent pol-
The pH starting value of the reaction medium was lutant, phenol, in the space of ca. 10 min. There was a
set to 3.5 because previous studies showed that this significant decrease in the total organic carbon content
was the optimal value for phenol oxidation by hydro- of the reaction medium. This indicated that most of
gen peroxide in the presence of iron [6,7]. Though the organic load was removed from the liquid-phase.
pH was left out of control during reaction, for all Deeper removal of the organic load was impeded, as
catalytic tests performed, no significant change was observed by other groups [6,7], by the accumulation
noted. of refractory organic byproduct acids that were sta-
Blank tests were preliminarily carried out prior to ble in the reaction medium because they were already
the actual heterogeneous catalytic tests, in the absence highly oxidized molecules.

100
90
Relative residual concentration

80 TOC
70 PhOH
60 H2O2

50
40
30
20
10
0
0 50 100 150
Time (min)

Fig. 1. Relative residual number of moles for TOC, phenol and H2 O2 . CWPO over B-c catalyst. Conditions: temperature 80 ◦ C; phenol
initial concentration 170 ppm C; 163% H2 O2 (excess stoichiometric ratio); catalyst load 0.5 g/l; Fe 25.3 mg.
298 N. Crowther, F. Larachi / Applied Catalysis B: Environmental 46 (2003) 293–305

Table 1 lishing that there was some parasitic self-degradation


Relative residual number of moles for TOC, phenol and H2 O2 of the hydrogen peroxide, especially in the early mo-
Catalyst Time Residual ψ (S in %) ments of the reaction where H2 O2 decomposition was
(min)
Phenol TOC H2 O2 at its highest. Similarly, the useful fraction of hydro-
(%) (%) (%) gen peroxide that contributed for the removal of the
A-c 10 2.2 69 58 5.3 (44) organic pollution (quantified in terms of a selectivity
20 0 48 47 4.0 (58) S = 2.33 × 100/ψ) indicates CWPO selectivity al-
60 0 24 28 3.7 (63) ways less than 100% (Table 1).
120 0 15 15 4.0 (58)
180 0 14 9 4.2 (56)
3.2. Catalyst stability to leaching and recycling, and
B-c 10 0.7 73 60 6.7 (35) mineralization selectivity
20 0 53 40 4.9 (48)
60 0 35 33 4.0 (58)
120 0 25 25 3.8 (61) To evaluate the contribution of homogeneous iron
180 0 19 19 3.9 (60) cations possibly leaching off from the solid catalysts,
C-c 10 0 58 60 3.6 (65)
an experiment in the same conditions was carried out
20 0 53 56 3.4 (69) using iron nitrate salts to mimic the leached cations.
60 0 46 41 3.6 (65) The iron concentration was chosen to be 5 ppm
150 0 35 28 3.8 (61) (1.5 mg/300 ml) which corresponded to the amount of
D-c 20 0.4 100 49 –
60 0 50 35 3.2 (73) Table 2
120 0 43 26 3.2 (73) Iron catalysts leaching patterns
150 0 37 24 3.0 (78) Catalyst Time Iron
(min)
D-e 20 0.4 86 76 6.0 (39) Initial Leacheda Max. leached
60 0 65 59 3.9 (60) (mg) (%) ± 0.5 conc. (ppm)b
130 0 60 54 3.9 (60) A-c 20 24.2 14 15
180 0 47 47 3.4 (69) 60 18
E-c 20 0 96 49 45.0 (0.05) 120 17
60 0 62 33 5.5 (42) 180 13
120 0 36 21 3.9 (60) B-c 20 25.3 2 5
180 0 27 11 3.8 (61) 60 5
Conditions: temperature 80 ◦ C, initial phenol: 180 ppm C, stoi- 120 5
chiometric ratio excess H2 O2 : 150%, catalyst load 0.5 g/l. 180 6
C-c 20 7.3 41 14
According to the mineralization reaction of phenol 60 54
by hydrogen peroxide: 150 56
D-c 20 2.2 64 5
C6 H5 OH + 14H2 O2 → 6CO2 + 17H2 O, 60 65
120 65
the molar ratio 150 66
n0H2 O2 − nH2 O2 (t) D-e 20 1.6 92 5
Ψ(t) = 60 100
n0TOC − nTOC (t) 130 100
180 84
measures the utilization of hydrogen peroxide ver-
sus the consumption of dissolved organic carbon as E-c 20 4.5 45 12
60 77
a function of time (Table 1). The ultimate value ex- 120 77
pected in the case of complete oxidation of phenol 180 75
and its intermediate organics and in the absence of
Conditions: see Table 1.
self-decomposition of the oxidant, would be 2.3. For a Leached amount is relative to the initial iron number of moles.
all catalysts, the values of ψ were over 2.3, estab- b Maximum amount of leached iron, expressed in mg/l.
N. Crowther, F. Larachi / Applied Catalysis B: Environmental 46 (2003) 293–305 299

100
90
Relative residual concentration

80 TOC
70 PhOH
60 H2O2

50
40
30
20
10
0
0 50 100 150

Time (min)

Fig. 2. Relative residual concentrations for TOC, phenol and H2 O2 . Homogeneously ferric catalyzed phenol wet peroxidation. Conditions:
temperature 80 ◦ C; phenol initial concentration 175 ppm C; 162% H2 O2 (excess stoichiometric ratio); Fe 1.5 mg.

Fe leached from pristine B-c catalyst (Table 2). The exhibited a lesser tendency to metal leaching. In par-
relative residual profiles for TOC, phenol and H2 O2 ticular, catalyst B-c was the one that exhibited the best
are illustrated in Fig. 2 for the homogeneously ferric performance in terms of leaching (6% or 5 ppm), be-
catalyzed wet peroxidation of phenol. Homogeneous sides being among the best to degrade phenol and to
iron cations, which are entirely available catalytic reduce TOC (Table 1).
centers, yielded a similar pattern of total phenol Although the expected maximum leaching would
oxidation and partial phenol mineralization. Phenol occur at the end of the reaction periods, materials A-c
vanished in about 25 min (Fig. 2) compared to the and D-e appeared to be an exception and showed max-
10 min required in the case of pristine B-c catalyst imum leaching during the course of CWPO. For sam-
(Fig. 1). In addition, after 150 min, ψ = 3.4 which ple D-e, this may be ascribed to the limited analytical
illustrates that here as well, the oxidant underwent
some self-degradation. 10
Summarized in Table 2 are the values of leached
iron amounts from the catalysts expressed in terms of 8
time course of Fe (%, w/w) with respect to the total
% Fe leaching

iron content of the pristine catalysts, and of the ulti- 6


mate ppm concentration in the solutions. Fig. 3 shows
the leached amount of iron versus time for catalyst
4
B-c, and was typical of the profiles monitored for the
other materials. The information in Table 2 evidences
2
that iron in all tested catalysts had a tendency to par-
tially dissolve in the liquid.
Two trends deserve attention at this stage. It can be 0
discerned that materials prepared without fluorides in 0 50 100 150
Time (min)
the synthesis gel, such as catalysts D and E, lost vir-
tually almost their whole iron inventory very quickly Fig. 3. Percent iron leaching for catalyst B-c. Conditions: temper-
(66–100% Fe leach off). Conversely, catalysts A–C, ature 80 ◦ C; phenol initial concentration 170 ppm C; 163% H2 O2
for which fluorides were added to the synthesis gel, (excess stoichiometric ratio); catalyst load 0.5 g/l; Fe 25.3 mg.
300 N. Crowther, F. Larachi / Applied Catalysis B: Environmental 46 (2003) 293–305

Table 3
Carbon content of the catalysts after reaction, including recycled catalysts
Catalyst Removed carbon form liquid (mg/l) Deposited carbon on solid (mg/l) Mineralization selectivity (%)

A-c – – –
B-c 125 8.1 94
B-c recycled 40 2.5 94
C-c 131 10.0 92
C-c recycled 86 2.7 97
D-c – – –
D-e – – –
E-c – – –
Removed carbon is given by TOC measurement, deposited carbon, expressed as equivalent of carbon concentration in the liquid, is given
by elemental analysis. Mineralization selectivity represents the amount of carbon actually converted to CO2 .

accuracy considering the lowest iron loading of this mass of carbon converted to CO2 divided by the mass
catalyst. At the other extreme, some re-adsorption of of carbon eliminated from water: [(TOC0 − TOC −
iron to the catalyst might be at the origin of the slight W)/(TOC0 − TOC)] × 100, where W is the concentra-
disturbance in the analysis of Fe from sample A-c. tion of “solid” carbon on the catalyst expressed as the
The catalysts before and after reaction were also mass of carbon per unit solution’s volume in mg/l.
investigated for their carbon content and thus for the As can be seen from Table 3, the mineralization se-
mineralization selectivity. The materials after initial lectivity for catalysts B-c, B-c-r, C-c and C-c-r was
activation did not contain any carbon. Table 3 presents very high suggesting that the amount of carbon sus-
the results of the carbon content detected in the mate- ceptible to build up on the catalyst was marginally low.
rials after the reaction runs. The removed carbon con- As no darkening of the materials was noticed during
centration corresponds to the difference between the the reaction, it is likely that catalyst deactivation, if it
initial (TOC0 ) and final TOC in the liquid. The de- was to occur, could not be ascribed to the deposition
posited carbon (W) corresponds to the carbon that left of carbonaceous materials.
the aqueous solution but that did not convert to CO2 . Table 4 shows the catalytic behavior of some reused
The catalyst mineralization selectivity is defined as the materials. All reused catalysts were calcined, just like

Table 4
Catalyst reuse after thermal reactivation
Recycled catalyst Time (min) Residual ψ Iron leaching

Phenol (%) TOC (%) H2 O2 (%) Initial (mg) Maximum

% ppm in liquid

A-c-r 10 94 97 100 – 10.7 36 13


60 0 53 45 3.9
120 0 46 36 4.1
B-c-r 10 99 100 86 High 6.5 2 0.5
30 73 92 82 11.1
120 4 69 49 7.1
C-c-r 10 96 91 94 2.4 1.0 30 0.9
60 0 50 48 3.9
140 0 40 37 3.9
D-c-r 20 85 100 99 – 0.1 100 0.3
50 78 91 91 3.7
120 68 82 86 3.1
Conditions: see Tables 1 and 2.
N. Crowther, F. Larachi / Applied Catalysis B: Environmental 46 (2003) 293–305 301

Table 5 peak around 2θ = 2◦ was the d100 lattice plane, and


Materials’ specific surface area and pore diameters was typical of hexagonal mesoporous structures [28]
Catalysts Specific surface Pore size (Fig. 4a–d).
area (m2 /g) distribution (Å) The template was removed either thermally or by
A-c 100 – solvent extraction. This process is expected to increase
C-c 480 33 the contrast of the X-ray diffraction since the pores are
D-c 1058 30
now empty. This was especially visible for catalysts C
D-e 866 30
E-c 895 – and D regardless of the activation method, revealing
hexagonal structure of the catalysts by emergence of
secondary diffraction peaks. Catalyst C was expected
to have very thick pore walls [17] and therefore the
for the activation, prior to recycling and contacting
secondary peaks were less obvious. In the case of sam-
with fresh phenolic solutions. It is noteworthy that the
ple E, only calcination revealed the hexagonal struc-
catalyst amount, phenol feed concentration as well as
ture, whereas extraction led to pore breakdown. Upon
temperature were held under the same conditions as in
calcination, samples A and B underwent, respectively,
the testing of the fresh catalysts. However there was no
partial and complete collapse of their structure. This
attempt to readjust the amount of catalyst to account
occurred for instance in lamellar mesoporous phases
for the loss of iron between the fresh and the recycled
such as MCM-50 type materials [28]. It is possible to
catalyst tests. For all the reused catalysts, a notice-
reckon that B exhibits lamellar symmetry, and that A
able activity loss was observed in comparison with the
displays a mixture of hexagonal (stable upon calcina-
activity using the pristine catalysts A-c through D-c.
tion) and lamellar structures.
Such activity loss in the case of spent sample D-c-r
XRD patterns were also taken after the first catalytic
was dramatic as phenol hardly converted and its min-
run. This revealed how the structure was affected by
eralization very limited. All other materials were ca-
the reaction conditions. For catalysts C–E, the hexag-
pable of performing, once more, complete elimination
onal network was retained, as evidenced by only lit-
of phenol as well as a notable TOC reduction.
tle changes in the respective XRD peaks’ height and
In terms of leaching during the second run, sample
shape. Yet for A, complete collapse was observed.
D-c-r lost completely its iron content. The poor cat-
BET measurements were performed on the fresh
alytic activity should here be attributed to the minute
and on the activated materials (Table 5). A high spe-
amounts of iron in sample D-c-r (0.1 mg) which can-
cific surface area is typical of mesoporous phases, es-
not sustain high conversion of the pollutant (32% phe-
pecially for well structured hexagonal phases such as
nol conversion and 18% TC conversion). In contrast,
sample D-c (see also XRD pattern). The pore diame-
sample B-c-r hardly lost any iron (2%) in the solution,
ter being controlled by the template, sizes of 30–33 Å
and its good robustness to retain iron in the second
correspond to the expected values for CTAB templat-
run is to be paralleled with the one it exhibited during
ing.
the first run (6% iron loss).
For sample A-c, no pore size distribution maxi-
mum could be established. Before calcination though,
3.3. Materials’ structural data a specific surface area of 261 m2 /g and a pore size
of 32 Å were determined. In a similar way, for sam-
The structural data of the solids used for this study ple B-c, BET measurement yielded 10 m2 /g, though
included low-angle X-ray diffraction patterns to check for the fresh uncalcined sample, a BET surface area
the crystalline phase (Fig. 4), nitrogen adsorption of 480 m2 /g and a pore size of ca. 30 Å were found.
isotherms (BET) for specific surface area and pore For these two materials, the results corroborated the
size (Table 4), and total iron content (Table 5). XRD observations, i.e. that the inner-porous network
The materials obtained after filtration of the synthe- collapsed during activation by calcination.
sis gel, i.e. before activation, are referred to as fresh. Table 6 presents the iron weight content of the ma-
At this point, the template was still trapped within the terials. For catalysts A and B, the iron loading in the
material. For all catalysts in the fresh form, the large calcined form was much higher than expected, which
302 N. Crowther, F. Larachi / Applied Catalysis B: Environmental 46 (2003) 293–305

Fig. 4. X-ray diffractograms for materials A–E, fresh, calcined, extracted, and after CWPO reaction for the initially calcined materials.

was rather surprising for silica-based materials. For sis of 6 g of silica (see experimental section above).
the others, the content after activation was close to the Preparations C–E yielded around 6 g of activated ma-
expected value. For all catalysts prepared in this work, terial, whereas samples A and B, prepared using a very
the amount of each reagent was determined on a ba- similar method, produced approximately 1.7 g [16]. It
N. Crowther, F. Larachi / Applied Catalysis B: Environmental 46 (2003) 293–305 303

Table 6 Sample E-c exhibited a slightly better catalytic ac-


Materials’ iron content (expressed as weight percentage) at differ- tivity than sample D-c or D-e. HMS-type materials,
ent steps: in the synthesis media, calculated from the amounts of
formed by using neutral templates as opposed to ionic
iron source added to the gel; before first use; before re-use
templates, presented better features for catalytic ap-
Catalysts In synthesis After activation After reusing plication than well defined hexagonal structures [29].
gel (%) (%) (%)
While catalyst D-e let all its iron go, catalyst
A 5.0 15.1 (c) 12.2
D-c could retain 33% after one cycle of utilization.
B 5.0 16.6 (c) 15.6
C 5.0 4.9 (c) 2.2 Moreover, preparatory work, involving other solvent-
D 2.5 1.5 (c) 0.5 extracted MCM-41 type mesoporous materials were
1.2 (e) – tested for CWPO, confirmed an almost total loss of
E 2.5 2.3 (c) – Fe. It may be argued that the solvent extraction of
the template, when compared to calcination, does not
bring better anchoring of iron into the structure. As
far as iron stability into the structure was concerned,
is likely that the precipitation of the silica from the catalysts D-c, D-e and E-c should be discarded.
sol–gel medium was incomplete, and thus the solid ac-
tually formed was richer in iron than the synthesis gel. 4.2. Catalysts A-c, B-c and C-c
After reusing the catalysts, the iron loading dropped
due to leaching during the CWPO reaction. Before the first run and after calcination, the meso-
porous structure was well defined for C-c, it was
partially altered for A-c but totally destroyed for
4. Discussion B-c. During CWPO reaction, A-c structure collapsed
completely, whereas that of C-c remained stable. The
This study’s goal was to investigate the potential stability of C-c can be explained by the wall thick-
of iron-containing mesoporous silicalite materials for ness obtained for synthesis gel pH around neutral,
the heterogeneous CWPO of phenol. For this purpose, as opposed to the alkaline conditions for A and B
two decisive factors need to be met, namely, the cat- samples.
alyst activity and the catalyst stability in the reaction Materials A-c and B-c were capable to achieve ca.
medium. Catalyst stability can be envisioned through 85% pollution removal which was equivalent or better
three criteria: deactivation by fouling, deactivation by than the TOC conversions reported in previous stud-
leaching, and structure robustness. Deactivation by ies in the case of Fe-HZSM-5 zeolite or pillared clays
fouling was ruled out based on the high mineraliza- [6,30]. All three materials A-c, B-c and C-c enabled
tion selectivities shown in Table 3. Iron leaching off very rapid and complete destruction of the parent pol-
was the main factor for deactivation for the five cat- lutant in the first run, and still exhibited good activity
alyst preparation protocols (Tables 2 and 4) although after reuse.
addition of fluorides in the synthesis gel appeared to C-c sample lost 56% of its iron in the first run leav-
lower the extent of leaching. ing it as less appealing than A-c and B-c samples with
less than 15% Fe loss. After recycling, C-c and A-c
4.1. Catalysts D-c, D-e and E-c iron loss amounted to a further 30–40% as opposed
to 2% Fe loss of B-c sample. The acidic environment
Catalysts D-c, D-e and E-c were particularly prone was probably a major factor for the extraction of the
to iron leach off, and respectively, 66, 100 and 75% of metal from the support, and B-c catalyst appeared to
the active species were lost during the first CWPO run. be acid-proof in terms of iron retention.
It was impossible to assess unambiguously the cat- Catalysts A-c, B-c and C-c were prepared from gels
alytic activity contributed by the heterogeneous iron. containing fluoride, introduced as hydrofluoric acid
Especially for samples D-e and E-c, which lost nearly (A and B) or as ammonium fluoride (C). They were
all their active sites, they could not be reused in a sec- intentionally used to favor iron insertion into the silica
ond run. matrix, so it was expected to be tightly bound to the
304 N. Crowther, F. Larachi / Applied Catalysis B: Environmental 46 (2003) 293–305

support. On the other hand, it is anticipated that the The structure of catalyst B-c was not mesoporous
extra-framework iron will leach off first. nor microporous, and exhibited a very low specific sur-
Catalyst B-c, freshly activated, was the solid that face area. Paradoxically, catalyst B-c displayed among
offered the best catalytic activity, in spite of the little the best activity and leach-proof ability compared to
estimated sites’ availability (low BET surface area). It any equivalent iron-modified mesoporous support. On-
was investigated by XRD for microporosity and local going work is in progress to investigate the nature of
crystalline environments provided by iron oxide clus- the catalytic sites on catalyst B. The question of the
ters, from 2θ diffraction angles of 1–60◦ but no mi- way iron is bound to the support, as framework (sub-
cropores or nanoclusters were detected. The change stituted into the silica matrix) or extra-framework (iron
in activity after recycling cannot be explained though nano-clusters trapped in the material) species could be
leaching deactivation due the low amounts of iron re- elucidated by ESR and Mössbauer spectroscopy cou-
leased in the solution. Besides, no change in color pled with EXAFS. Also, the availability of iron to-
was noticed. It is presumed that a structural change wards the incoming oxidant or organic molecules may
occurred during the first and/or second runs that in- be clarified by TPR, TPO, TPD or TPA.
volved the masking of some active centers without
migration of iron from framework to extra-framework
positions. Acknowledgements
The leaching propensity of sample A-c in the two
runs was obvious. This may be correlated to the Financial support from the Natural Sciences and En-
fact that during calcination of spent catalyst, prior to gineering Research Council of Canada (NSERC) and
re-use, turned the material to dark brown color and the Fonds Québécois de la Recherche sur la Nature
involved migration of Fe from framework to extra- et les Technologies (FQRNT) is gratefully acknowl-
framework (i.e. exposed) positions prompting leach edged.
off of iron in the course of the CWPO reaction.
Sample C-c was likely to exhibit better catalytic
activity because of well defined mesopores, thick References
walls and high sites availability, compared to the
collapsed porous networks in A-c and B-c samples. [1] J. Levec, Chem. Biochem. Eng. Q. 11 (1998) 47.
[2] Y.I. Matatov-Meytal, M. Sheintuch, Ind. Eng. Chem. Res. 37
It was noticed, however, that the color turned pale
(1998) 309.
during CWPO but darkened upon activation cycles. [3] S. Imamura, Ind. Eng. Chem. Res. 38 (1999) 1749.
Migration of iron as in sample A-c might occur as a [4] Z.Y. Ding, M.A. Frisch, L. Li, E.F. Gloyna, Ind. Eng. Chem.
consequence of calcination. Res. 35 (1996) 3257.
[5] R. Wandeler, A. Baiker, CATTECH 4 (2000) 128.
[6] J. Barrault, M. Abdellaoui, C. Bouchoule, A. Majesté, J.M.
Tatibouët, A. Louloudi, N. Papayannakos, N.H. Gangas, Appl.
5. Conclusion
Catal. B: Environ. 27 (2000) 225.
[7] K. Fajerwerg, J.N. Foussard, A. Perraed, H. Debellefontaine,
The best candidate for heterogeneous catalytic wet Water Sci. Technol. 35 (4) (1997) 103.
peroxide oxidation of phenol, under mild conditions [8] M. Falcon, B. Peyrille, P. Reilhac, J.-N. Foussard, H.
(atmospheric pressure and 80 ◦ C), would be the cal- Debellefontaine, Rev. Sci. de l’Eau 6 (1993) 411.
[9] M. Falcon, K. Fajerweg, J.-N. Foussard, E. Peuch-Costes,
cined catalyst B. In the space of 15 min, the parent
M.T. Maurette, H. Debellefontaine, Environ. Technol. 16
pollutant was totally degraded and after 180 min, 81% (1995) 501.
of the organic load was removed from the liquid-phase [10] L. Plant, M. Jeff, Chem. Eng. (September) (1994) EE16.
with a mineralization selectivity of 93%. It was also [11] S. Biz, M.L. Occedi, Catal. Rev.-Sci. Eng. 40 (1998) 329.
possible to reuse B-c after re-calcination, although it [12] S. Hamoudi, F. Larachi, A. Sayari, Catal. Lett. 77 (2001) 227.
[13] A. Katovic, G. Giordano, B. Bonelli, B. Onida, E. Garrone,
showed less catalytic activity. In terms of robustness,
P. Lentz, J.B. Nagy, Micropor. Mesopor. Mater. 44–45 (2001)
B-c endured leaching of only 6% after the first cat- 275.
alytic run, and when re-calcined for re-use, of only [14] S.E. Dapurkar, S.K. Badamali, P. Selvam, Catal. Today
2%. 68 (1–3) (2001) 63.
N. Crowther, F. Larachi / Applied Catalysis B: Environmental 46 (2003) 293–305 305

[15] Y. Wang, Q. Zhang, T. Shishido, K. Takehira, J. Catal. 209 (1) [23] M. Strlic, J. Kolar, B. Pihlar, Acta. Chim. Slov. 46 (4) (1999)
(2002) 186. 555.
[16] M. Alves, H.O. Pastore, Micropor. Mesopor. Mater. 47 (2001) [24] I.W.C.E. Arends, R.A. Sheldon, Appl. Catal. A: Gen. 212
397. (2001) 175.
[17] L. Pasqua, F. Testa, R. Aiello, F. Di Renzo, F. Fajula, [25] C.T. Kresge, M.E. Leonowicz, W.J. Vartuli, J.S. Beck, Nature
Micropor. Mesopor. Mater. 44–45 (2001) 111. 359 (1992) 710.
[18] Z. Gabelica, S. Valange, Res. Chem. Intermed. 24 (3) (1998) [26] K.M. Reddy, C. Song, Catal. Lett. 36 (1999) 103.
227. [27] A. Tuel, S. Gontier, Chem. Mater. 8 (1996) 114.
[19] J.W. Hill, R.H. Petrucci, General Chemistry, second ed., [28] J.S. Beck, J.C. Vartuli, W.J. Roth, C.T. Kresge, M.E.
Prentice-Hall, Englewood Cliffs, NJ, 1999. Leonowicz, K.D. Schmitt, C.T.-U. Chu, D.H. Olsen, E.W.
[20] S. Hamoudi, K. Belkacemi, F. Larachi, Chem. Eng. Sci. 54 Sheppard, S.B. McCullen, J.B. Higgins, J.L. Schlenker, J.
(1999) 3569. Am. Chem. Soc. 114 (1992) 10834.
[21] S.T. Hussain, A. Sayari, F. Larachi, Appl. Catal. B: Environ. [29] P.T. Tanev, T.J. Pinnavaia, Nature 368 (1994) 321.
34 (2001) 1. [30] K. Fajerwerg, H. Debellefontaine, Appl. Catal. B 10 (1996)
[22] S.T. Hussain, A. Sayari, F. Larachi, J. Catal. 201 (2001) 153. 229.

You might also like