Sunteți pe pagina 1din 9

Applied Catalysis A: General 206 (2001) 221229

Iron molybdate catalysts for methanol to formaldehyde oxidation:


effects of Mo excess on catalytic behaviour
A.P.V. Soares a , M. Farinha Portela a, , A. Kiennemann b , L. Hilaire b , J.M.M. Millet c
a GRECAT Grupo de Estudos de Catlise Heterognea, Instituto Superior Tcnico,
Lisbon Technical University, Avenida Rovisco Pais, 1049-001 Lisbon, Portugal
b ECPM LERCSI, UMR CNRS 7515, 25, Rue Becquerel, BP08, 67087 Strasbourg Cedex 2, France
c Institut de Recherches sur la Catalyse, CNRS, 2 Avenue Albert Einstein, 69626 Villeurbanne Cedex, France
Received 10 June 1999; received in revised form 28 April 2000; accepted 28 April 2000

Abstract
Two iron molybdate catalysts have been prepared by coprecipitation, one stoichiometric (Mo/Fe atomic ratio=1.5) and the
other with a typical industrial composition (Mo/Fe atomic ratio=3). Physicochemical characterisation shows that Mo excess
brings about an increase in the surface area and deconvolution of NH3 TPD curves evidences some changes in acidity. On
the other hand, surprisingly, Mo excess does not cause significant changes in the activity for methanol oxidation per unit
surface area. This provides evidence that Fe2 (MoO4 )3 is in fact the active phase of the catalyst. The catalyst with Mo excess
leads to higher selectivity for HCHO to the detriment of dimethyl ether selectivity. This result is attributable to larger oxygen
availability at the catalyst surface and changes in acidity. The catalytic behaviour of the prepared catalysts was compared with
that of an industrial catalyst. It was found that the prepared catalyst with an Mo/Fe composition similar to the industrial one
is more active due to larger surface area. 2001 Elsevier Science B.V. All rights reserved.
Keywords: Iron molybdates; Methanol oxidation; Formaldehyde

1. Introduction Nowadays, both processes are still in use [3] and


the choice between silver and iron molybdate cata-
Formaldehyde is one of the most important basic lysts must be based not only on economic aspects
chemicals and is required for the manufacture of a but should also take into account the product end-use,
large number of industrial and consumer products. size of plant and type of operation [4]. In comparison
It is the most important industrially produced alde- with catalysis over Ag, the oxidation of methanol over
hyde [1]. Two processes are generally used in the in- Fe2 O3 MoO3 is carried out at lower temperatures, the
dustry to produce formaldehyde, both using methanol catalyst is less sensitive to contamination by normal
as the starting material [2]: (i) dehydrogenation of methanol impurities and also provides 95% selectivity
methanol-rich air mixture over silver catalysts and (ii) to formaldehyde [1,2].
direct oxidation of methanol-poor air mixture over iron Since the pioneering work of Adkins and Peter-
molybdate catalysts. son [5], the growing industrial importance of FeMo
oxide catalysts has led to a large number of inves-
Corresponding author. Fax: +351-21-8477695. tigations, but important aspects of their catalytic
E-mail address: mportela@alfa.ist.utl.pt (M. Farinha Portela). behaviour are still controversial. Several researchers

0926-860X/01/$ see front matter 2001 Elsevier Science B.V. All rights reserved.
PII: S 0 9 2 6 - 8 6 0 X ( 0 0 ) 0 0 6 0 0 - 1
222 A.P.V. Soares et al. / Applied Catalysis A: General 206 (2001) 221229

have found that the catalytic behaviour of iron molyb- and Mo-rich (Mo/Fe=3) catalysts were compared in
dates is mainly dependent on their Mo/Fe atomic this work.
ratio [613]. MoO3 exhibits high selectivity (nearly
100%), but has poor activity, whereas Fe2 O3 has
poor selectivity for HCHO. The addition of Fe2 O3 to 2. Experimental
MoO3 leads to an increase in activity, proportional
to the iron content, up to an Mo/Fe atomic ratio=1.7 2.1. Preparation of catalysts
[7]. Further enrichment with iron leads to progres-
sive decrease in activity. Normal iron molybdate
Iron molybdates with Mo/Fe atomic ratios of 1.5
(Fe2 (MoO4 )3 ) is likely to be the active component
(stoichiometric mixed oxide) and 3 (industrial-like
of the catalyst [7,12,14]. However, some researchers
catalyst) were prepared by coprecipitation from an
associate active sites with surface Mo atoms in oc-
aqueous solution of Fe(NO3 )3 9H2 O and (NH4 )6 Mo7
tahedral coordination and such coordination is only
O24 . Iron nitrate solution was slowly added to the
achieved in Fe-defective iron molybdates [10,13].
ammonium heptamolybdate solution under vigorous
This result agrees with the fact that maximum activity
stirring. For preparation of the stoichiometric mixed
is found for catalysts with an Mo/Fe ratio greater than
oxide, Mo solution was first acidified (pH2) with
the stoichiometric value. The presence of two termi-
HNO3 . After total addition of Fe solution, the precip-
nal oxygen atoms double-bonded to Mo in such a
itates were ripened in contact with mother liquors at
coordination allows the reacting methanol molecules
373 K for 3 h under stirring. During this period, the
to be bonded simultaneously at two points to the
Mo/Fe=1.5 yellowish precipitate changes in colour
surface. The hydrogen abstraction of the methanol
to pale green, whereas the Mo/Fe=3 precipitate only
hydroxyl group produces metoxy species that are in-
undergoes a slight intensification of its initial yellow
termediates in formaldehyde formation. The role of
colour.
Fe in iron molybdate catalysts would be to favour
After ripening, the precipitates were filtered, dried
the transfer of O2 and H2 O between the surface and
at 393 K overnight, and finally calcined. Calcination
the gas phase and to facilitate reoxidation of reduced
was performed at 648 K for 10 h in air flow.
Mo [13,15]. Furthermore, Fagherazzi and Pernicone
[8] admitted that the presence of Fe3+ ions increases
the concentration of methanol adsorption sites, con- 2.2. Characterisation of catalysts
sisting of an anion vacancy (acid sites) and an O2
(basic sites). The BET surface areas were measured with a
According to Alessandrini et al. [14], the active Perkin-Elmer Shell 212 C sorptometer by N2 ph-
phase of iron molybdate catalysts is the stoichiomet- ysisorption. The bulk elemental compositions were
ric mixed oxide. However, these researchers recognise determined by atomic absorption. X-ray diffrac-
the need for Mo excess to maintain the structural co- tion (XRD) patterns were recorded with a D5000
herence of the catalyst, to increase the surface area Siemens diffractometer with a primary beam quartz
and to prevent the formation of iron rich phases during monochromator (Co K1 =1.78897 ) at 40 kV and
reaction. 25 mA. The morphology, chemical analysis and ho-
Several studies on catalyst deactivation [1620] mogeneity of the prepared catalysts were examined
show that, during reaction, the catalyst loses Mo and with a scanning electron microscope JEOL JSM 840
Mo excess is required to maintain the Mo-rich com- with a Delta Kevex energy dispersive X-ray analyser.
position. This would be the reason why industrial cat- FT-IR spectra were recorded with a Perkin-Elmer
alysts always have Mo/Fe atomic ratios greater than 1600 spectrometer in the range of 4000400 cm1 ,
the stoichiometric value. In fact, an Mo/Fe atomic with a resolution of 4 cm1 . For each analysis, 1 mg
ratio=3 is frequent in industrial catalysts [2]. of the product was ground with 100 mg of KBr
In order to clarify the role of Mo excess in catalytic and then pelleted (2000 kg cm2 ) to a disc 13 mm
systems based on MoFe mixed oxides for methanol in diameter. Sixty-four scans were used for each
selective oxidation to formaldehyde, stoichiometric spectrum.
A.P.V. Soares et al. / Applied Catalysis A: General 206 (2001) 221229 223

X-ray photoelectron spectroscopy (XPS) was used (30.0 cm3 min1 ). After reaching 723 K, the sample
to investigate the chemical analysis and reduction state was cooled to 323 K and ammonia adsorption was
of the surface of the fresh catalysts. The analyses performed by NH3 pulses (1 cm3 ). The sample was
were performed with a VG ESCA 3 apparatus. The then flushed with He for 2 h to remove ammonia ex-
kinetic energy of the emitted photoelectron was given cess. Ammonia was thermally desorbed (5 K min1 )
by EKin =1486.6EBcor , where 1486.6 eV is the en- in an He flow from 323 to 723 K. The amount of
ergy of the incident radiation (Al K) and EBcor the desorbed NH3 was measured by TCD detector.
corrected binding energy of the electron. The binding
energy of the electron, corrected from charge effects, 2.3. Catalytic tests
was calibrated with respect to the signal for adventi-
tious carbon (binding energy: 284.8 eV). The corrected Methanol oxidation was carried out in a con-
intensity for an element A is given by [21]: ventional flow apparatus at atmospheric pressure.
Catalytic behaviour was studied in steady-state con-
IAmeas ditions. Feed mixtures were prepared by injecting
IAcor = p
nA A EKinA liquid methanol into air flow with a precise Gilson
302 pump. The catalyst was diluted with inert SiC
where IAmeas is the measured intensity (experimen- (1:7 mass ratio), to avoid adverse thermal effects, and
tal), nA the scan number (experimental), and A the charged into a tubular Pyrex reactor with thermocou-
cross-section (tabulated [22]). The ratio of the cor- ple in a coaxial centred thermowell. The reactor outlet
rected intensities of two elements A and B directly was kept at 403 K, to prevent condensation of liquid
accounts for the A/B atomic ratio at the surface of the products and formaldehyde polymerisation, and was
analysed sample. Curve fitting analysis was performed connected to a Shimadzu GC-8A gas chromatograph
using a DoniachSunjic line shape with no asymmetry with a thermal conductivity detector. Two in-parallel
parameter and a Shirley background correction. Due columns, one at room temperature (Hayesep D) for
to the small energy separation between Fe2+ and Fe3+ N2 +O2 , CO and CO2 , and the other in the oven
species, the spectra of pure reference compounds con- of the chromatograph (porapak T) at 443 K for
taining either Fe2+ or Fe3+ were recorded under the HCHO, CH3 OCH3 (DME), H2 O, CH3 OH (MeOH),
same conditions. In addition to the binding energies, HCOOCH3 (MF) and CH2 (OCH3 )2 (DMM), allowed
the parameters, which represent the half width at half the separation of the reactor effluents. Formic acid
maximum of the peaks, were extracted from reference was never detected.
spectra. Their values were used in the curve fitting
procedure for the catalysts and it was obvious that the
parameters of two species, and not one, had to be in- 3. Results and discussion
troduced to get a reasonable fit. Moreover, the binding
energies found in such a way were in good agreement, 3.1. Characterisation of catalysts
within 0.2 eV, with those of reference samples.
The bulk composition of iron phases was studied The results presented in Table 1 show that the sur-
by Mssbauer spectroscopy at room temperature with face area of the prepared catalysts increases with Mo
a time mode spectrometer and a constant acceleration content in the bulk. For the same Mo/Fe atomic ra-
drive. 57 Co in an Rh matrix was used as a radiation tio, the industrial catalyst has a lower surface area
source. The isomer shifts () were given with respect
to -Fe and calculated, as the quadrupolar splittings Table 1
BET surface areas
(1) and the line width (W) with a precision of about
0.02 mm s1 . Catalyst Mo/Fe Surface
(atomic ratio) (atomic ratio) area (m2 g1 )
The study of the surface acidity of the prepared
catalysts was performed by ammonia TPD. The Industrial 3 5
Coprecipitated 3 8.3
catalyst sample was previously submitted to a pre- Coprecipitated 1.5 2.8
treatment involving heating (5 K min1 ) under He
224 A.P.V. Soares et al. / Applied Catalysis A: General 206 (2001) 221229

Table 2
Mo and Fe composition and surface reduction of studied fresh
catalysts
Catalyst (preparation Mo/Fe (atomic ratio) Fe2+ /Fe3+
Mo/Fe atomic ratio) (atomic ratio)
Atomic EDX XPS XPS
absorption
1.5 1.5 2.1 1.89 0.24
3 2.9 4.3 2.35 0.19
Industrial 3.1

than the prepared catalyst, which can be attributed to


a more severe thermal treatment during the activation
step (773 K instead of 648 K).
Several authors have prepared iron molybdates with
different Mo/Fe atomic ratios, but the study of the
variation of their surface areas as a function of the
ironmolybdenum content has never been reported.
However, Sun-Kuo et al. [13] present, for equivalent
Mo/Fe compositions, higher surface areas than those
found in the present work. This disagreement can be
attributed to a shorter thermal treatment: 2 h instead
of the 10 h used here.
The results of the bulk and surface chemical anal-
ysis of the catalysts are summarised in Table 2.
The Mo/Fe bulk ratios of the coprecipitated solids
determined by atomic absorption are in good agree- Fig. 1. Scanning electron micrographs of prepared catalysts.
ment with those of the parent solutions used for the
precipitations. The noted differences are within the
experimental error range. EDX analysis shows an Mo ratio and pH during the precipitation step and the age-
enrichment of the surface of the prepared catalysts. ing period. But the conditions reported by the authors
The same is observed by XPS for a stoichiometric to favour the formation of ordered lamellae structure
catalyst but not for the Mo/Fe=3 catalyst. It is well (low pH and Mo/Fe atomic ratios with long ageing pe-
known that XPS is more surface-sensitive than EDX riods) correspond to those used for the precipitation of
and a profile of Mo contents would be established in the stoichiometric phase in the present work. In con-
the catalyst particles by the balance of the rates of Mo clusion, the morphology exhibited by this last phase
migration from the bulk and Mo loss from the surface. is in accordance with the results of these authors.
The scanning electron micrographs (Fig. 1) show No fibre-like material, usually assignable to the
that the catalysts with stoichiometric composition dis- MoO3 phase [13,18], was found in the prepared
play an ordered lamellae morphology, whereas cata- catalysts, which is a good indication that no phase
lysts with Mo excess show a sponge-like morphology. segregation occurred during the preparation and cal-
The morphological appearance is thus well in accor- cination steps. The segregation of MoO3 is a frequent
dance with surface areas of the catalysts. However, and adverse occurrence during the preparation of iron
this finding seems to disagree with recent results of molybdenum mixed oxides. In fact, such an occur-
Sun-Kuo et al. [13] who found that samples with or- rence must be prevented because the MoO3 phase
dered lamellae have larger surface areas and that the is much less active for the selective oxidation of
morphology of iron molybdates depends on the Mo/Fe methanol than iron molybdate. Moreover, segregation
A.P.V. Soares et al. / Applied Catalysis A: General 206 (2001) 221229 225

Table 3
X-ray diffraction (XRD) results of the prepared catalysts
Mo/Fe=1.5 Mo/Fe=3

d () I/I0 (%) d () I/I0 (%)

3.884 100 3.248 100


3.481 38.6 3.854 85.4
3.583 29.5 3.447 71.3
4.098 23.3 3.791 45.0
4.357 22.2 2.306 33.7
3.249 22.2 2.649 33.0
3.207 21.6 4.064 26.4
2.966 19.3 1.845 25.2
2.629 18.8 6.935 23.8
6.436 16.5 3.549 20.5
5.809 14.2 2.948 18.8

in the 700900 cm1 range can be ascribed to tetra-


hedric species of Mo in Fe2 (MoO4 )3 [10].
Iron-containing phases were also analysed by
Mssbauer spectroscopy. The significant spectral
parameters of industrial and prepared catalysts are
Fig. 2. X-ray diffraction (XRD) patterns of prepared catalysts: (a)
Mo/Fe atomic ratio=1.5 and (b) Mo/Fe atomic ratio=3. summarised in Table 4. The values were obtained by
fitting the experimental spectra with pure Lorentzian

of MoO3 can lead to an Fe-rich phase that promotes


total oxidation of methanol [6].
Stoichiometric mixed oxide displays a higher crys-
tallinity than mixed oxides with Mo excess. This
result agrees with the fact that Mo excess retards the
crystallisation of Fe2 (MoO4 )3 [7]. X-ray patterns of
fresh catalysts presented in Fig. 2 confirm that the sto-
ichiometric phase is better crystallised than the phase
with Mo excess. The indexed XRD patterns (Table 3)
show that the Mo/Fe=1.5 catalyst corresponds only
to the orthorhombic ferric molybdate [23], while the
Mo/Fe=3 catalyst presents, in addition, MoO3 (peaks
at 3.248 and 6.935 ) [10]. The absence of MoO3
peaks in the Mo/Fe=1.5 X-ray patterns does not totally
rule out the existence of free MoO3 because Deltch-
eff et al. [24] have demonstrated that MoO3 is only
detected by XRD when its content is higher than 7%.
Mo excess in the prepared Mo/Fe=3 catalyst and
industrial catalyst were confirmed by the presence in
IR spectra (Fig. 3) of a narrow band for all catalysts
at 990 cm1 and a broad one at 624 cm1 character-
istic of MoO3 [10,13,25]. The weak and narrow band
that appears at 960 cm1 can be assigned to FeOMo Fig. 3. IR spectra of catalysts: (a) Mo/Fe atomic ratio=1.5; (b)
bond vibrations [25]. Finally, the broad band observed Mo/Fe atomic ratio=3; (c) industrial catalyst.
226 A.P.V. Soares et al. / Applied Catalysis A: General 206 (2001) 221229

Table 4
Mssbauer parameters of fresh catalystsa
Catalyst Fe species (mm s1 ) W (mm s1 ) 1 (mm s1 ) Relative intensity (%) Phase attribution

Industrial Fe3+ 0.41 0.32 0.21 100 Fe2 (MoO4 )3


Mo/Fe=3 Fe3+ 0.44 0.26 0.19 99 Fe2 (MoO4 )3
Fe2+ 0.89 0.28 2.34 <1
Mo/Fe=1.5 Fe3+ 0.44 0.27 0.20 100 Fe2 (MoO4 )3
a : isomeric shift; W: band width; 1: quadrupolar splitting.

curves. The identification of the phases was based on stoichiometric catalyst. The Mo/Fe=1.5 catalyst has
the published results of previous investigations of iron a broader desorption peak of chemisorbed NH3 at the
molybdates [26,27,33]. The results show that iron in lower temperature. For the strongest acid sites, the
all catalysts is present only in the Fe2 (MoO4 )3 phase. peak was higher for the Mo/Fe=3 catalyst (Fig. 5).
Mssbauer parameters of ferric cations in the cat-
alysts with Mo excesses are comparable to those of 3.2. Catalytic behaviour
ferric cation in the stoichiometric catalyst. This result
shows well that MoO3 excess does not form a solid Only formaldehyde, DME, MF, H2 O and small
solution with the Fe2 (MoO4 )3 phase, which is in amounts of CO2 were found as reaction products in
agreement with the results of Abaulina et al. [10]. The all catalytic tests.
prepared catalyst with Mo excess additionally presents In Fig. 6, conversion of methanol versus reaction
a small amount of ferrous cations (<1%) that cannot temperature is plotted for the same contact time for
be identified with ferrous cations in - or -FeMoO4 . all catalysts. As expected, conversion increases with
The surface reduction state was checked by XPS. temperature and this effect seems to be more intense
Iron-reduced species were found in the studied cata- at low conversions. On a weight basis, it is seen that
lysts. The results in Table 2 show a noticeable sur- the industrial catalyst and the Mo/Fe=3 catalyst are
face reduction. This result agrees with the Mssbauer much more active than the Mo/Fe=1.5 one (Fig. 6a),
findings. Mo excess seems to make the iron reduction
difficult.
Characterisation of the surface acidity of the pre-
pared catalysts has been carried out by NH3 TPD.
Both the stoichiometric catalyst and the catalyst pre-
pared with Mo excess display a desorption spectrum
with two distinct desorption peaks (Fig. 4). Both pre-
sented thermograms were normalised to the same sur-
face area. Deconvolution of experimental curves by
symmetrical Gaussian curves [28] reveals a third des-
orption peak for both catalysts (Fig. 5). Stoichiometric
mixed oxide display desorption peaks with maxima
at 400, 520 and 600 K, whereas for the Mo/Fe=3
catalyst, the maxima are at 400, 480 and 580 K. For
both catalysts, the low-temperature peak has its max-
imum at the same temperature. It can be assigned to
physisorbed NH3 . The Mo/Fe=3 catalyst physisorbs
more NH3 than the stoichiometric one. The other des-
orption maxima of chemisorbed NH3 are at lower tem-
peratures for Mo/Fe=3 catalysts than for Mo/Fe=1.5 Fig. 4. TPD of NH3 : () Mo/Fe atomic ratio=3; () Mo/Fe
ones. This means that acid sites are stronger for the atomic ratio=1.5. Lines correspond to the deconvoluted curves.
A.P.V. Soares et al. / Applied Catalysis A: General 206 (2001) 221229 227

Fig. 5. Deconvolution of experimental TPD of NH3 curves with symmetrical Gaussian curves.

but on a surface basis, they are comparable (Fig. 6b). our finding is that, due to the rate of loss of Mo from
The results show that the surface Mo/Fe atomic ra- the surface, the composition and structure of the top
tio does not significantly affect methanol conversion layers of the catalyst with Mo excess are identical to
since the specific activity of catalysts is almost in- the ones of the stoichiometric molybdate. The data of
dependent of the Mo content. This finding disagrees Table 2 do not disagree with such a hypothesis taking
with the majority of previously published results that into account that XPS is sensitive to the composition
reported an optimum activity for a catalyst composi- in a layer of a few nanometres.
tion corresponding to an Mo/Fe atomic ratio of 1.7 In Fig. 7 are plotted changes in the products
[7], but is consonant with those of Alessandrini et al. selectivities with the reaction temperature. As re-
[14] that reported no effect of Mo excess on the spe- ported previously [29,30], formaldehyde selectivity
cific activity of iron molybdates. One explanation for increases with methanol conversion, except for a

Fig. 6. Oxidation of methanol in an air feed with 4.4% of methanol. Steady-state conversion vs. temperature at the same contact time: (a)
weight basis W/F=4.9 gcat h mol1 2 1
MeOH ; (b) surface area basis S/F=40.7 mcat h molMeOH .
228 A.P.V. Soares et al. / Applied Catalysis A: General 206 (2001) 221229

high conversion level when the consecutive formalde-


hyde oxidation is not negligible. At a low conversion
level, catalysts with Mo excess are more selective for
HCHO than stoichiometric ones that produce high
amounts of DME, even at low contact time [31]. In
principle, this result would be explainable on the
basis of acidity differences. In fact, it is well known
that the formation of ethers occurs by an SN 2 mecha-
nism with one molecule of the alcohol acting as the
nucleophile and with the other, protonated molecule
of the alcohol acting as the substrate. NH3 TPD for
both catalysts shows differences in the position of the
central desorption peak at around 500 K and peak ar-
eas. For the stoichiometric catalyst, the maximum of
the peak is positioned at 520 K instead of 480 K for
the Fe-defective catalyst and its area its also larger,
evidencing stronger acidity of such a type. However,
Edwards et al. [32] have ascribed the high forma-
tion of DME to the low availability of oxygen at the
catalyst surface. Furthermore, Ai [33,34] stated that
dehydration of methanol to ether occurs over acidic
oxides with no oxidising potential. This would allow
to conclude that Mo in excess increases the amount
of available oxygen at the surface and enhances the
oxidation potential of mixed oxide.

4. Conclusions

The present investigation has shown that the intro-


duction of Mo excess in MoFe mixed oxide, with
respect to stoichiometry, leads to an increase in the
surface area of the catalysts.
The characterisation results agree with the litera-
ture pointing out that Mo in excess is embedded in the
interstices of the crystalline network without substitu-
tion of Fe3+ ions by Mo6+ ions.
In spite of the acidic character of the Mo oxide,
the prepared catalysts (Mo/Fe=1.5 and Mo/Fe=3) do
not present very significant differences of acidity, as
determined by NH3 TPD.
The results of the catalytic tests evidence that the
intrinsic activity (per unit surface area) is not af-
fected by Mo excess, suggesting that the Fe2 (MoO4 )3
Fig. 7. Oxidation of methanol: steady-state conversion and selec- stoichiometric phase would be the active phase
tivities (sel.) vs. temperatures (4.4% of methanol in air feed and of the catalyst. However, Mo excess increases the
W/F=4.9 gcat h mol1 MeOH ). HCHO selectivity to the detriment of DME selec-
tivity. The differences in DME selectivity found
A.P.V. Soares et al. / Applied Catalysis A: General 206 (2001) 221229 229

with stoichiometric and Mo-rich catalysts can be [13] M.R. Sun-Kou, S. Mendioroz, J.L.G. Fierro, J.M. Palacios,
attributable to differences in acidity and oxidising po- A. Guerrero-Ruz, J. Mater. Sci. 30 (1995) 496.
tential. Mo excess in bulk seems to increase surface [14] G. Alessandrini, L. Cairati, P. Forzatti, P.L. Villa, F. Trifir,
J. Less-Comm. Met. 54 (1977) 373.
reoxidability. [15] J. Novakova, P. Jiru, J. Catal. 27 (1972) 155.
[16] I. Popov, V.N. Bibin, G.K. Boreskov, Kinet. Catal. (Engl.
Transl.) 17 (2) (1976) 322.
References [17] N. Burriesci, F. Garbassi, M. Petrera, S. Petrini, in: B.
Delmon, P. Grange, P. Jacobs, G. Poncelet (Eds.), Catalyst
[1] G. Reuss, W. Disteldorf, O. Grundler, A. Hilt, in: Ullmanns Deactivation, Elsevier, Amsterdam, 1979, p. 279.
Encyclopedia of Industrial Chemistry, 5th Edition, Vol. A11, [18] J. Arruano, S. Wanke, Can. J. Chem. Eng. 53 (1975) 301.
VCH, Weinheim, 1992, p. 619. [19] Y.H. Ma, S.J. Kmiotek, J. Catal. 109 (1988) 132.
[2] B. Stiles, T.A. Koch, in: Catalyst Manufacture, 2nd Edition, [20] N. Pernicone, Catal. Today 11 (1991) 85.
Marcel Dekker, New York, 1995, Chapter 20, p. 197. [21] M.B. Ward, M.J. Lin, J.H. Lundsford, J. Catal. 50 (1977) 306.
[3] ECN Process Review, April 1994, p. 30. [22] J.H. Scofield, J. Electron Spectrosc. Relat. Phenom. 8 (1976)
[4] R. Chauvel, P.R. Curty, R. Maux, C. Petitpas, Hydrocarbon 129.
Process. 52 (9) (1973) 179. [23] V. Massarotti, G. Flor, A. Marini, J. Appl. Cryst. 14 (1981)
[5] H. Adkins, W.R. Peterson, J. Am. Chem. Soc. 53 (1931) 64.
1512. [24] R. Deltcheff, M. Amirouche, M. Che, J.-M. Tatibout, M.
[6] G.D. Kolovertnov, G.K. Boreskov, V.A. Dzisko, B.I. Popov, Fournier, J. Catal. 125 (1990) 292.
D.V. Tarasova, G.C. Belugina, Kinet. Catal. (Engl. Transl.) [25] A.B. Anagha, S. Ayyappan, A.V. Ramaswamy, J. Chem. Tech.
6 (6) (1965) 950. Biotechnol. 59 (1994) 395.
[7] G.K. Boreskov, G.D. Kolovertnov, L.M. Kefeli, L.M. [26] G. Petrini, F. Garbassi, M. Petrera, N. Pernicone, in:
Plyasova, L.G. Karakchiev, V.N. Mastikhin, V.I. Popov, V.A. H.F. Barry, P.C. Mitchell (Eds.), Chemistry and Uses of
DzisKo, D.V. Tarasova, Kinet. Catal. (Engl. Transl.) 7 (1) Molybdenum, Climax Molybdenum Company, Ann Arbor,
(1965) 125. MI, USA, 1982, p. 437.
[8] G. Fagherazzi, N. Pernicone, J. Catal. 16 (1970) 321. [27] A.W. Sleight, B.L. Chamberland, J.F. Weiher, Inorg. Chem.
[9] J.-M. Leroy, S. Peirs, G. Tridot, Comptes Rendus Acad. Sci. 7 (6) (1968) 1093.
Paris, series C Janvier (1971) 218. [28] J.Y. Neira, J.J. Godoy, R. Cid, Bol. Soc. Chil. Qum. 36
[10] L.I. Abaulina, G.N. Kustova, R.F. Klevtsova, B.I. Popov, V.N. (1991) 141.
Bibin, V.A. Melekhina, V.N. Kolomiichuk, G.K. Boreskov, [29] H.-Y. Chen, Mater. Res. Bull. 14 (1979) 1583.
Kinet. Catal. (Engl. Transl.) 17 (5) (1976) 1126. [30] W.-H. Cheng, J. Catal. 158 (1996) 477.
[11] Yu.V. Maksimov, M.Sh. Zurmuktashvili, I.P. Suzdalev, L.Ya. [31] A.P. Vieira Soares, Ph.D. Thesis, Technical University of
Margolis, V. Krylov, Kinet. Catal. (Engl. Transl.) 25 (4) (1984) Lisbon, 1996.
809. [32] J. Edwards, J. Nicolaidis, M.B. Cutlip, C.O. Bennett, J. Catal.
[12] V. Demidov, I.G. Danilova, G.N. Kustova, L.M. Plyasova, 50 (1977) 24.
N.G. Skomorokhova, L.L. Sedova, V.B. Nakrokhin, B.I. [33] M. Ai, J. Catal. 52 (1978) 16.
Popov, Kinet. Catal. (Engl. Transl.) 33 (1992) 910. [34] M. Ai, J. Catal. 54 (1978) 426.

S-ar putea să vă placă și