You are on page 1of 12

Applied Catalysis B: Environmental 43 (2003) 81–92

Perovskite-type oxides synthesized by reactive grinding


Part IV. Catalytic properties of LaCo1−x Fex O3
in methane oxidation
V. Szabo a , M. Bassir a , A. Van Neste b,c , S. Kaliaguine a,∗
a Department of Chemical Engineering, Laval University, Ste Foy, Que., Canada G1K 7P4
b Department of Mining and Metallurgy, Laval University, Ste Foy, Que., Canada G1K 7P4
c Nanox Inc., 335 St. Joseph E., Quebec, Que., Canada G1K 3B3

Received 27 March 2002; received in revised form 29 October 2002; accepted 31 October 2002

Abstract
In this part of the present work, the catalytic properties in methane oxidation of perovskite-type mixed-oxides, LaCo1−x Fex O3
prepared by reactive grinding of the component oxides were studied. A fine control of the nature and concentration of the
various sites active in oxido-reduction catalysis was performed by applying the comprehensive picture of the effects of ther-
mal pretreatment of the mixed-oxides generated by this technique, developed in part I of this contribution. These properties
which were previously illustrated by the kinetics of complete oxidation of n-hexane over five nanocrystalline LaCo1−x Fex O3
catalysts treated in parts II and III are now demonstrated by a kinetic study of methane oxidation over the same catalysts.
© 2002 Elsevier Science B.V. All rights reserved.
Keywords: Perovskites; Methane oxidation; Kinetics; Reactive grinding

1. Introduction of perovskites for oxidation is a proven technology that


has reached commercial status in some instances [15].
The interest for atmosphere VOC emission preven- The main limitations of the perovskite-type materials
tion and clean-up augmented much during the last are their low surface area typically around 20 m2 /g
decades in relation with the increasingly strict legis- [15,16]. Various preparation methods were proposed
lations [1,2]. A variety of catalysts are used for this in order to increase the specific surface area (SSA)
purpose and in general precious metals are preferred and, thus catalytic activity [15,17]. In the last decade,
as total oxidation catalysts [3–6]. Transition metal ox- mechanical alloying by high-energy ball-milling was
ides are also utilized for this purpose and in particular, developed to produce novel oxide materials [18], in-
the perovskite-type oxides are considered as poten- cluding nanophase of fluorite-structured CeO2 -ZrO2
tially significant [7–14]. These materials are cheaper, catalysts and iron carbides catalysts synthesis [19,20].
showing thermal stability in atmospheres containing This grinding method yielded gains in catalytic ac-
oxygen and steam at high temperature (gas turbine, tivity when compared with traditional syntheses
catalytic burners) and resistance to poisoning. The use [21–23] and was shown to be appropriate for prepar-
ing mixed-oxide catalysts with redox properties [19].
∗ Corresponding author. Moreover a beneficial effect of ball-milling would
E-mail address: kaliagui@gch.ulaval.ca (S. Kaliaguine). be to increase specific surface area [23]. Reactive

0926-3373/02/$ – see front matter © 2002 Elsevier Science B.V. All rights reserved.
doi:10.1016/S0926-3373(02)00308-9
82 V. Szabo et al. / Applied Catalysis B: Environmental 43 (2003) 81–92

grinding, a method which was demonstrated recently increase the specific surface area. The powder quantity
in our laboratory, allows to increase SSA between 3 obtained for each sample was 5 g.
and 150 m2 /g in a controlled manner [24,25]. Sample 4 was prepared in a larger tempered steel
Methane can be taken as a reference compound vial (183.5 cm3 ) with nine balls made of the same
because it belongs to the most difficult to oxidize material. The same SPEX ball mill was rotated at
class of VOCs [26]. As a consequence in the present 700 rpm. The grinding lasted for 28 h in an oxygen at-
study, a detailed kinetic analysis of the catalytic mosphere with additive. The perovskite quantity syn-
oxidation of methane on five LaCo1−x Fex O3 per- thesized was 15 g.
ovskite catalysts was performed. In reporting this For each preparation, the milling proceeds at room
work, the parameters of the kinetic model are dis- temperature. When the synthesis used an additive, the
cussed in relation with the results reported in parts I, perovskite end product was thoroughly washed, in or-
II and III of this contribution [25,27,28]. Part I deals der to free the sample from even traces of this additive.
with a temperature-programmed desorption of oxy-
gen (TPDO) and X-ray photoelectron spectroscopy 2.2. Catalyst characterization
(XPS) of LaCo1−x Fex O3 materials prepared by reac-
tive grinding. Parts II and III report a kinetic analysis 2.2.1. Powder XRD
of the oxidation of n-hexane over the same materi- Conversion into perovskite upon grinding and the
als. The methane oxidation study presented in this nature of product phases were determined by XRD us-
fourth part comprises an experimental analysis of the ing Cu K␣ radiation (λ = 1.54184 Å, Siemens D5000
poisoning effects of water vapor and sulfur dioxide. or Philips diffractometer). The conditions of spectral
recording were in a 2θ angle step scanning from 10 to
80◦ , with a 2.4 s delay for each 0.05◦ step.
2. Experimental procedures
2.2.2. Nitrogen adsorption
2.1. Preparation of LaCo1−x Fex O3 The specific surface area, at different calcination
temperatures, for each LaCo1−x Fex O3 sample was
Five LaCo1−x Fex O3 , with different specific surface measured by nitrogen adsorption (BET method) at
areas, were prepared by reactive grinding. Powders −196 ◦ C using an Omnisorb Sorptometer 100 instru-
of lanthanum oxide, La2 O3 (Alfa 99.99%), cobalt ox- ment operated in the continuous mode. Samples of
ide, Co3 O4 (Baker & Adamson, 97.49%) and Fe2 O3 about 1 g were outgassed at 200 ◦ C for approximately
(Strem Chemicals, 99.8%), in stoichiometric propor- 24 h under vacuum before starting the adsorption–
tions were ball-milled at high-energy. The La2 O3 desorption experiments. N2 adsorption measurements
powder was preliminarily calcined at 600 ◦ C for 24 h, were performed up to a relative pressure P/P0 = 0.1.
which makes it possible to eliminate any trace of lan-
thanum hydroxide. Perovskite sample 1 was prepared 2.2.3. Atomic absorption
in a tungsten carbide (WC) vial (39.4 cm3 ) with three The chemical composition analyses (La, Co, Fe) of
WC balls, rotated at 700 rpm using a SPEX labora- the perovskite end products and the residual impurities
tory grinder (SPEX is a trademark of SPEX Indus- were performed by atomic absorption spectroscopy
tries, Edison, NJ). The vial atmosphere was air. The using a Perkin-Elmer 1100B spectrometer. Samples
milling duration was 12 h and the perovskite quantity were dissolved in 10% HCl solution at 60 ◦ C.
obtained, 5 g. These conditions are not optimized to In part I of this contribution, detailed studies of
obtain a high specific surface area. these catalysts by XPS and TPDO were reported [25].
Samples 2, 3 and 5 were synthesized in a tempered These studies have shown that the temperature of cal-
steel (TS) vial (64.6 cm3 ) with three TS balls, using the cination of LaCo1−x Fex O3 perovskites produced by
SPEX grinder rotated at 1100 rpm. Their preparation reactive grinding is a key factor in determining the na-
was carried out over 24 h in an air atmosphere for 2 and ture and the density of oxygen adsorption sites as well
3 and in an oxygen atmosphere for 5. As discussed in as their energy of adsorption. Interactions with molec-
[25] additives were introduced in the mill in order to ular oxygen during calcination stabilize the surface
V. Szabo et al. / Applied Catalysis B: Environmental 43 (2003) 81–92 83

oxygen vacancies responsible for adsorption of ␣ oxy- tector and two HayeSep DB (2 × 5 m, i.d. 1.0 mm)
gens and the doubly charged vacancies (Co2+ 䊐 Co2+ ) columns.
which are the desorption sites for ␤ oxygens during
the TPDO experiment. ␤ oxygens originate from the 2.4. Data treatment
bulk of the LaCoO3 lattice.
The same procedure as in the second and third parts
2.3. Oxidation reaction of this contribution was used to perform the data treat-
ment. Methane steady state conversion results (main-
The methane oxidation reaction was carried out in tained for at least 2 h) at the outlet of the reactor were
a tubular fixed bed micro-reactor with a quartz reactor collected for the five LaCox Fe1−x O3 catalysts at dif-
tube (7 mm i.d.) at atmospheric pressure. A gas mix- ferent flowrates over a large range of temperature. It
ture of oxygen, neon and methane (2.5% CH4 , 87.5% was verified that the catalytic activity was not affected
O2 , 10% Ne, Praxair) and pure oxygen gas (100% by the series of tests provided the reaction temperature
oxygen, Praxair) were used to prepare the feed mix- does not exceed the calcination temperature. Because
ture of 0.25% CH4 , 98.75% O2 and 1% Ne, using the reaction was performed at higher temperature than
two mass flow controllers (Omega 0152 E, Matheson n-hexane oxidation [27,28] the kinetic data could be
Model 8240). In these conditions, the adiabatic tem- represented using the simplified kinetic equation pro-
perature rise at 100% CH4 conversion was calculated posed by Arai et al [29]. The mechanism consists in
to be 20 ◦ C. only one surface Mars–Van Krevelen process involv-
The feed mixture stream passed through a catalyst ing ␤ oxygens.
charge of about 100 mg, installed in the reactor and Using the same procedure as in parts II and III of this
secured by using plugs of quartz wool. The reactor contribution, rate values were obtained by first cross-
was placed in a furnace equipped with a temperature plotting experimental conversions X as a function of
programmer. The reaction temperature was controlled W/F (where W is the weight of the catalyst and F the
with a K-type thermocouple placed inside the catalytic inlet molar flowrate of methane) at several selected
bed. The total flow rate of the gas mixture was in reaction temperatures. The resulting curves were thus
the range of 7.5–40 ml/min, resulting in a range of fitted using the following exponential expression:
gas hourly space velocity of 5625–30,000 h−1 (STP    
W
ml/h ml catalyst). Before the catalytic tests the samples X = a 1 − exp −b (1)
F
were treated in He for 1 h at ambient temperature.
When the influence of water vapor partial pressure The derivatization of Eq. (1) led to obtaining point
on reaction rate was studied, the same gas mixture values for the rate of reaction:
  
of oxygen, neon and methane as above was bubbled dX W
through a water containing saturator maintained at a r= = ab exp −b (2)
d(W/F) F
constant temperature between 33.1 and 54.3 ◦ C. This
allowed obtaining mole fractions of water vapor be- where r represents the reaction rate in the conditions
tween 5 and 15% and the methane content was main- prevailing at the outlet of the catalyst bed. Note that
tained at 0.25%. Eqs. (1) and (2) are not kinetic equations but were
For the SO2 poisoning study, the mixture of oxygen, fitting of conversion data allowing to established
neon and methane (2.5% CH4 , 87.5% O2 , 10% Ne, meaningful numerical estimations of reaction rates in
Praxair) was blended with another SO2 containing gas various conditions.
mixture (26 ppm SO2 , 10% Ne, balance oxygen, Prax- Then the values of r were represented using Arai
air) using the two mass flow controllers (Omega 0152 et al. simplified equation [29]:
E, Matheson Model 8240) in order to obtain 0.25% r = kPCH4 (3)
CH4 and 23 ppm SO2 .
Analyses of reactants and products were performed This kinetic model (Eq. (3)) involves only the partici-
using a gas chromatograph (GC, Hewlett Packard HP pation of one site (␤ oxygens) in the catalytic process.
6890 Series) equipped with a thermal conductivity de- In parts II and III for n-hexane total oxidation which
84 V. Szabo et al. / Applied Catalysis B: Environmental 43 (2003) 81–92

was performed at lower temperature [27,28] two kinds Fig. 1B yielding a set of isothermal curves for methane
of oxygen species (␣ and ␤) coexisted at the surface conversion as a function of the pseudo-contact time
of these LaCo1−x Fex O3 perovskite-type oxides. At W/F. Then, these isothermal curves are fitted with ex-
the temperatures where CH4 oxidation was performed ponential Eq. (1). Numerical estimates of the local re-
the coverage of ␣ sites, and therefore the correspond- action rate (2) were obtained by derivatizing of these
ing oxidation rates, are negligible. Thus the methane fitting lines analytically. The numerical estimates of r
oxidation reaction is believed to involve a Mars–Van were fitted to Eq. (3) by linear regression [30] and the
Krevelen mechanism using oxygen atoms originating lines shown in Fig. 1A were obtained by numerical
from the bulk. These are identical with the oxygen integration over the length of the reactor
atoms that desorb as ␤ molecular oxygen in TPDO  XE
W dX
experiments. = (4)
F 0 r
assuming isothermal reaction. The hypothesis of an
3. Experimental results isothermal reactor was thought acceptable as the dif-
ference in measured temperatures at bed inlet and
The X-ray diffraction patterns of the five samples outlet was found to be lower than 5 ◦ C at complete
(1–5) were reported in part III of this communica- conversion. The observed agreement between the ex-
tion [28]. These samples which were studied here for perimental X data and calculated curves in Fig. 1A,
their activity in methane oxidation were essentially is a test of validity for the values of the rate constant
crystallographically pure nanocrystalline perovskites k obtained by the fitting procedure.
LaCo1−x Fex O3 . The crystal size as determined by Figs. 2–5 show the methane conversion in catalytic
X-ray line broadening was indeed found to be in the tests of catalyst 2 (Fig. 2), catalyst 3 calcined at 500 ◦ C
9.5–16.5 nm range. (Fig. 3), catalyst 1 calcined at 700 ◦ C (Fig. 4) and
Table 1 gives values of the BET specific surface area catalyst 5 calcined at 600 ◦ C (Fig. 5). The Arrhenius
measured after calcination at various temperatures and plots for the rate constant k in Eq. (3) are shown in
the elemental composition of these five samples. It was Fig. 6:
verified that these calcinations had no effect on the k = k0 e−E/RT (5)
XRD pattern of the solids. In this work, the calcination
temperatures were higher than in the n-hexane study The pre-exponential factors (k0 ) and activation ener-
reported in part III [28] due to the higher reaction gies (E) were obtained by linear regression of these
temperatures reached in methane oxidation. lines. These constants are recapitulated in Table 1. The
Fig. 1A shows the methane conversion data obtained data for sample 1 reported in Table 1 and Fig. 4 were
at steady state for various reactor temperatures and obtained after calcination of this sample to 700 ◦ C. In
at various GHSV, using catalyst 4 calcined at 600 ◦ C. parts II and III of this work, the sample 1 was only
Complete oxidation was reached between the temper- calcined at 400 ◦ C which was sufficient to obtain some
atures of 470–500 ◦ C, which shows the high activity of activity in n-hexane oxidation over ␣ oxygen sites.
these materials. The data in Fig. 1A are cross-plotted in Sample 1 (designated as N1 in [25]) calcined at this

Table 1
Kinetic parameters established for the different samples
Samplea Elemental composition Calcination temperature (◦ C) Specific surface area (m2 /g) EA (kcal/mol) k0 (mol/s g atm)

1 La1.0 Co0.96 Fe0.003 O3−∆ 700 2.4 24.5 1.97 × 102


5 La0.99 Fe1.01 O3−∆ 600 13 22.4 3.14 × 102
2 La0.88 Co0.95 Fe0.17 O3−∆ 500 21.4 25.5 5.08 × 103
3 La0.91 Co0.92 Fe0.17 O3−∆ 500 29 25.5 4.28 × 103
4 La0.96 Co0.7 Fe0.34 O3−∆ 600 37.8 26.3 1.32 × 104
a According to the notation used in parts II and III of this communication [27,28].
V. Szabo et al. / Applied Catalysis B: Environmental 43 (2003) 81–92 85

Fig. 1. (A) Methane oxidation conversion over sample 4 calcined at 600 ◦ C: (䊉) 5625 h−1 , (䊏) 11,250 h−1 , (䉱) 16,875 h−1 , (䉲) 22,500 h−1 ,
(䊊) 30,000 h−1 . Reaction conditions: 0.25% CH4 , 98.75% O2 , 1% Ne and 100 mg of catalyst. The lines are calculated by integrating
Eq. (3) over the length of the reactor. (B) Values of methane conversion over sample 4 calcined at 600 ◦ C as function of pseudo-contact
time (W/F). Points are experimental data, lines are calculated using Eq. (1): () 355 ◦ C, (䉭) 375 ◦ C, (䊐) 394 ◦ C, (䊊) 412 ◦ C, (䉲) 442 ◦ C,
(䉱) 471 ◦ C, (䊏) 486 ◦ C, (䊉) 504 ◦ C.

temperature showed no activity in methane oxidation. Methane oxidation rate values obtained using cata-
A similar situation was met with sample 3, which dis- lyst 3 calcined at 500 ◦ C derived from the conversion
played no activity in methane oxidation, when cal- data reported in Fig. 3, are reported in Fig. 7. In this
cined at 300 ◦ C. figure, reaction rates at given temperatures are plotted
86 V. Szabo et al. / Applied Catalysis B: Environmental 43 (2003) 81–92

Fig. 2. Methane oxidation conversion over catalyst 2 calcined at 500 ◦ C: (䊉) 5625 h−1 , (䊏) 11,250 h−1 , (䉱) 16,875 h−1 , (䉲) 22,500 h−1 ,
(䊊) 30,000 h−1 . Reaction conditions: 0.25% CH4 , 98.75% O2 , 1% Ne and 100 mg of catalyst. The lines are calculated by integrating
Eq. (3) over the length of the reactor.

Fig. 3. Methane oxidation conversion over catalyst 3 calcined at 500 ◦ C: (䊉) 5625 h−1 , (䊏) 11,250 h−1 , (䉱) 16,875 h−1 , (䉲) 22,500 h−1 ,
(䊊) 30,000 h−1 . Reaction conditions: 0.25% CH4 , 98.75% O2 , 1% Ne and 100 mg of catalyst. The lines are calculated by integrating
Eq. (3) over the length of the reactor.
V. Szabo et al. / Applied Catalysis B: Environmental 43 (2003) 81–92 87

Fig. 4. Methane oxidation conversion over sample 1 calcined at 700 ◦ C: (䊉) 5625 h−1 , (䊏) 11,250 h−1 , (䉱) 16,875 h−1 , (䉲) 22,500 h−1 ,
(䊊) 30,000 h−1 . Reaction conditions: 0.25% CH4 , 98.75% O2 , 1% Ne and 100 mg of catalyst. The lines are calculated by integrating
Eq. (3) over the length of the reactor.

Fig. 5. Methane oxidation conversion over sample 5 calcined at 600 ◦ C: (䊉) 5625 h−1 , (䊏) 11,250 h−1 , (䉱) 16,875 h−1 , (䉲) 22,500 h−1 ,
(䊊) 30,000 h−1 . Reaction conditions: 0.25% CH4 , 98.75% O2 , 1% Ne and 100 mg of catalyst. The lines are calculated by integrating
Eq. (3) over the length of the reactor.
88 V. Szabo et al. / Applied Catalysis B: Environmental 43 (2003) 81–92

Fig. 6. Arrhenius plots of constant k in Eq. (3) for the different catalysts.

Fig. 7. Linearity test for the rate of methane oxidation over catalyst 3 (designated as V1 in [25]) calcined at 500 ◦ C.
V. Szabo et al. / Applied Catalysis B: Environmental 43 (2003) 81–92 89

against the partial pressure of methane as a test of and 3–500) whereas the least active sample 1 catalyst
Eq. (3). yields clearly lower values of k and k2 . This similarity
seems to confirm the hypothesis that these two reac-
tions occur on the same sites.
4. Kinetic analysis and discussion The data for k are considered more reliable due to
the lower number of adjustable constants in the rate
The confirmation that the rate data follow Eq. (3) equation (Eq. (3)) for methane oxidation.
allows to conclude that on these catalysts, the active In Table 1, the catalysts are in increasing order of
oxygen species for methane total combustion is ␤ oxy- specific surface area and it is seen that this is also
gen with rate r (or r2 in parts II and III [27,28]). This essentially the increasing order of k0 values. Since
is in agreement with the results in the TPDO and O 1s the activation energy is almost the same for all cata-
XPS studies reported in the first part of this article [25] lysts (with the exception of sample 5) this is also the
which showed that ␣ oxygens are in very low surface order of increasing reaction rate. The specific surface
concentrations for samples calcined at temperatures area is therefore a very significant factor determining
exceeding 500 ◦ C. Fig. 8 allows a comparison between the rate of methane oxidation. Obviously, the effects
the rate constants k in Eq. (3) and k2 the rate constant of the composition and calcination temperature are
derived in [28] for the specific rate of n-hexane oxi- also significant. In particular, catalyst 5 having no
dation over ␤ oxygens. In spite of the different tem- Co (LaFeO3 ) is clearly less active than the cobalt
perature ranges at which the two processes occur, an containing catalysts.
obvious similarity is observed between the two rate It is interesting to note in Fig. 6 that the data ob-
constants. Not only the energies of activation E and tained with catalyst 5 which contains no Co show
E2 appear essentially equal (24.6 ± 1.1 kcal/mol for a sudden change in slope as temperature exceeds
E2 compared to 25.8 ± 0.3 kcal/mol for E2 in Co con- 530 ◦ C. Such a drop in the slope of the Arrhenius plot
taining catalysts), but very close values are observed is usually indicating the onset of a regime where in-
in both cases for the most active catalysts (2, 3–300 ternal diffusion becomes rate limiting. It is probably

Fig. 8. Comparison between rate constants k and k2 of the two reactions: methane oxidation and n-hexane oxidation respectively (k2 values
are from [28]).
90 V. Szabo et al. / Applied Catalysis B: Environmental 43 (2003) 81–92

meaningful that this behavior is obtained with the all should not be compared with the values for the other
Fe perovskite. It may be that the diffusional process catalysts.
which becomes rate limiting is the outward diffu- Fig. 10 shows the influence of water vapor at dif-
sion of oxygen ions O2− involved in the Mars–Van ferent partial pressures (5, 10 and 15% of water) for
Krevelen process. The results in Fig. 6 therefore sug- methane oxidation at 11,250 h−1 over catalyst 3 cal-
gest that the mobility of the oxygen ions is lower in cined at 500 ◦ C. The methane conversion is essentially
LaFeO3 than in the cobalt containing perovskites. not further affected as the partial pressure of water
Fig. 9 represents the values obtained for k0 /SSA exceeds 0.05 atm. This result seems to show that al-
which pictures the specific reaction rate per unit sur- ready for a partial pressure of 5%, perovskite surface
face area, as a function of the catalyst calcination sites are saturated. Moreover, after each experiment
temperature. It is interesting to note that no rate is ob- performed in the presence of water vapor, the catalyst
served at calcination temperatures lower than 500 ◦ C is calcined again at 500 ◦ C for 2 h in air, the initial ac-
which is exactly the range of calcination temperature tivity measured in absence of water vapor is restored.
at which TPDO experiments [25] indicated no forma- The deactivation by H2 O is therefore reversible.
tion of the surface sites responsible for the release of Fig. 11 was obtained with a catalyst produced in
␤ oxygens. Then increasing the calcination temper- conditions similar to catalyst 1 except that the calci-
atures up to 600 ◦ C yields a continuous increase in nation temperature was 500 ◦ C (compared to 700 ◦ C
reaction rate. This would correspond to an increased for catalyst 1) and that the specific surface area was
surface concentration of these sites due to an increased slightly higher (3.5 m2 /g). This Figure compares the
surface Co2+ stabilization process by interaction with methane conversion obtained as a function of time
oxygen during calcination as determined by XPS and at 500 ◦ C and GHSV of 5625 h−1 in two situations.
TPDO (see the mechanism described in [25]). At cal- Curve a was obtained with the standard feed (0.25%
cination temperatures above 600 ◦ C, the defect healing CH4 , 1% Ne, balance O2 ) and shows no deactiva-
process starts diminishing the surface defects con- tion over 3 days. Curve b was for a similar experi-
centration leading to a decreased methane oxidation ment with 23 ppm SO2 in the feed. It shows that the
rate. LaCoO3 catalyst is not resistant to sulfatation. Cur-
The data point for catalyst 5 is not shown in Fig. 9. rent work in our laboratory is examining the possibil-
As the activation energy E is different for this cat- ity to substitute La by a cation which yields less stable
alyst (see Table 1), the pre-exponential factor k0 sulfates.

Fig. 9. Effect of the calcination temperature on the specific frequency factor for methane oxidation.
V. Szabo et al. / Applied Catalysis B: Environmental 43 (2003) 81–92 91

Fig. 10. Study of the influence of PH2 O at 11,250 h−1 over catalyst 3 calcined at 500 ◦ C: (䊉) without water, (䊊) 5% H2 O, (䉲) 10%
H2 O, () 15% H2 O and (䊐) after catalyst regeneration at 500 ◦ C.

Fig. 11. Effect of SO2 poisoning over a LaCoO3 perovskite at 5625 h−1 and 500 ◦ C: (䊏) without SO2 , (䊉) with 23 ppm SO2 .

5. Conclusions the kinetic parameters of the various materials syn-


thesized. In particular, it was shown in part I of this
This work makes it possible to relate the prepara- contribution that controlling the calcination tempera-
tion and pre-treatment conditions of nanocrystalline ture allows to modify the surface concentration, the
LaCo1−x Fex O3 perovskites obtained by reactive heat of oxygen adsorption, and the nature of a vari-
grinding to their catalytic activity in methane oxida- ety of surface defects associated with two types of
tion. A detailed kinetic analysis allowed to compare adsorbed oxygens. It was proposed in parts II and III
92 V. Szabo et al. / Applied Catalysis B: Environmental 43 (2003) 81–92

that these two surface oxygens are ion radicals O2 − [9] K.S. Song, D. Klvana, J. Kirchnerova, Appl. Catal. A: Gen.
and O− . The respective roles of these two species in 213 (1) (2001) 113.
the oxidation of n-hexane was clarified in part III by [10] D. Kießling, R. Schneider, P. Kraak, M. Haftendorn, G.
Wendt, Appl. Catal. B: Environ. 19 (1998) 143.
means of a detailed kinetic study. In the present paper [11] G. Sinquin, J.P. Hindermann, C. Petit, A. Kiennemann, Catal.
it was established that owing to the higher reaction Today 54 (1) (1999) 107.
temperature of methane oxidation, only the O− ion [12] L.G. Tejuca, J.L.G. Fierro (Eds.), Properties and Applications
radical formed at the surface sites which allow the of Perovskite-Type Oxides, Marcel Dekker, New York, 1992.
thermal desorption of ␤ oxygens are utilized. This [13] Y.Ng. Lee, R.M. Lago, J.L.G. Fierro, V. Cortés, F. Sapiña,
E. Martı́nez, Appl. Catal. A: Gen. 207 (2001) 17.
was indeed indicated both by the simple rate Eq. (3) [14] H. Falcón, M.J. Martı́nez-Lope, J.A. Alonso, J.L.G. Fierro,
which reflects the Mars–Van Krevelen mechanism Solid State Ionics 131 (2000) 237.
and by the fact that no catalytic activity was observed [15] E.A. Lombardo, M.A. Ulla, Res. Chem. Intermed. 24 (5)
for catalysts treated in conditions which do not allow (1998) 581.
the formation of the surface defects associated with [16] V. Cortés Corberán, J. Roman. Catal. Soc. 6 (2) (1997) 113.
[17] R.J. Bell, G.J. Millar, J. Drennan, Solid State Ionics 131
the evolution of ␤ oxygens in TPDO experiments. (2000) 211.
The present part IV confirms that reactive grinding [18] C.J.H. Jacobsen, J. Jiang, S. Morup, B.S. Clausen, H. Topsoe,
is a perovskite synthesis technique which permits to Catal. Lett. 61 (3–4) (1999) 115.
reach high catalytic activities in oxidation reactions, [19] A. Trovarelli, F. Zamar, J. Llorca, C. de Leitenburg, G.
not only by enhancing the specific surface area but Dolcetti, J.T. Kiss, J. Catal. 169 (1997) 490.
[20] A. Trovarelli, P. Matteazzi, G. Dolcetti, A. Lutman, F. Miani,
also by controlling the surface density of specific sur- Appl. Catal. A: Gen. 95 (1993) L9.
face defects through a proper choice of calcination [21] S. Mori, W.-C. Xu, T. Ishidzuki, N. Ogasawara, J. Imai, K.
conditions. Kobayashi, Appl. Catal. A: Gen. 137 (1996) 255.
[22] V.A. Zazhigalov, J. Haber, J. Stoch, L.V. Bogutskaya, I.V.
Bacherikova, Appl. Catal. A: Gen. 135 (1996) 155.
References [23] F.C. Jentoft, H. Schmelz, H. Knözinger, Appl. Catal. A: Gen.
161 (1997) 167.
[1] W.J. Cooper, L.A. Holloway, Nature 384 (1996) 313. [24] S. Kaliaguine, A. Van Neste, US Patent 6,017,504 (2000).
[2] M. Amann, M. Lutz, J. Hazard. Mater. 78 (1–3) (2000) [25] S. Kaliaguine, A. Van Neste, V. Szabo, J.E. Gallot, M. Bassir,
41. R. Muzychuk, Appl. Catal. A: Gen. 209 (2001) 345.
[3] H. Widjaja, K. Sekizawa, K. Eguchi, H. Arai, Catal. Today [26] H. Taylor, R. O’Leary, Appl. Catal. B: Environ. 25 (2000)
47 (1999) 95. 137.
[4] H. Windawi, Z.C. Zhang, Catal. Today 30 (1996) 99. [27] V. Szabo, M. Bassir, A. Van Neste, S. Kaliaguine, Appl.
[5] D. Roth, P. Gelin, M. Primet, E. Tena, Appl. Catal. A: Gen. Catal. B: Environ. 37 (2) (2002) 175.
203 (1) (2000) 37. [28] V. Szabo, M. Bassir, J.E. Gallot, A. Van Neste, S. Kaliaguine,
[6] S. Minicò, S. Scirè, C. Crisafulli, R. Maggiore, S. Galvagno, Appl. Catal. B: Environ., in press.
Appl. Catal. B: Environ. 28 (2000) 245. [29] H. Arai, T. Yamada, K. Eguchi, T. Seiyama, Appl. Catal. 26
[7] P. Ciambelli, S. Cimino, S. De Rossi, L. Lisi, G. Minelli, P. (1986) 265.
Porta, G. Russo, Appl. Catal. B: Environ. 29 (2000) 239. [30] J.H. Seinfeld, Lapidus, Mathematical Methods in Chemical
[8] L. Marchetti, L. Forni, Appl. Catal. B: Environ. 15 (1998) Engineering, vol. 3: Process Modeling, Estimation and Identi-
179. fication, Prentice-Hall, Englewood Cliffs, NJ, 1978.

You might also like