You are on page 1of 12

Applied Catalysis, 74 (1991) 125-136 125

Elsevier Science Publishers B.V., Amsterdam

Temperature-programmed reduction and zeta potential


studies of the structure of MoO,/Al,O, and MoOJSiO,
catalysts
Effect of the impregnation pH and molybdenum loading

R. Lopez Corder0
Centro de Znvestigaciones Quimicas, Washington 169, La Habana (Cuba)

F.J. Gil Llambias


Departamento de Quimica, Facultad de Ciencia, Universidad de Santiago de Chile, Casilla
5659, Santiago (Chile), tel. (+56-2)6811644, fax. (+56-2)6812108

and
A. Lopez Agudo*
Znstituto de Catalisis y Petroleoquimica, C.S.Z.C, Serrano 119,28006 Madrid (Spain), tel. (+34-
1)2619400, fax. (+34-l )5642431

(Received 18 January 1991, revised manuscript received 18 March 1991)

Abstract

The influence of the impregnation pH (at 2-11) and molybdenum content (6-18 wt.-% MOO,) on
the surface structure of Mo03/A1203 and MoO,/SiO, catalysts has been studied by temperature-pro-
grammed reduction, X-ray diffraction and zeta potential measurements. The results indicate that mol-
ybdena was relatively well dispersed on A&O3 support, the dispersion being worse for the catalysts pre-
pared in acidic medium than for those in basic medium; formation of a crystalline phase was observed
only at a pH of 2 and for MOO, loadings > 14 wt.-%. However, on SiOz support only crystalline MOO,
species were found, irrespective of the preparation conditions. In this case the temperature-programmed
reduction profiles suggest the presence of two MoOB bulk phases which may probably differ in crystal
size, and whose proportion depends on the pH of the impregnating solution.

Keywords: catalyst preparation (wet impregnation), molybdenum/alumina, molybdenum/silica, pH


effect, temperature-programmed reduction, zeta potential.

INTRODUCTION

The nature and surface distribution of the molybdenum species present on


supported molybdena catalysts, which depend strongly on the preparation con-
ditions (impregnation procedure, molybdenum loading, pH of impregnation,
drying and calcination conditions, nature of the support, etc. ) , have been the

0166-9834/91/$03.50 0 1991 Elsevier Science Publishers B.V.


126

subject of numerous investigations using various physico-chemical techniques


[l-6 and references therein], including temperature-programmed reduction
(TPR). This technique has provided very useful information on the nature
and the strength of the interaction between the molybdenum species and sup-
port [ 781. Recently, the sensitivity of the TPR technique to detect the changes
in the surface structure of alumina-supported molybdena catalysts produced
by phosphorus incorporation has been shown [ 9,101. Since the phosphorus is
usually added as phosphoric acid, which implies the use of impregnating solu-
tions with different pH, depending on the preparation method and the loadings
of phosphorus and active metal, it is thus interesting to determine the influ-
ence of the pH of the impregnating solution, independently of the effect of
phosphorus.
The effect of the impregnation pH on the adsorption of the molybdenum
species and the final structure of the MOO, catalysts has been examined in
various studies [2-6,111, generally using Raman spectroscopy and, recently,
molybdenum-95 NMR spectroscopy [6,11]. In spite of the experimental sim-
plicity of the TPR technique it has not been used to examine systematically
the effect of the impregnation pH on the structure of Mo03/A1203 catalysts. A
recent work [ 121 deals specifically with the influence of the impregnation pH
on the MoOJSiO, system, but it was limited to a fixed loading of 8 wt.-%
MOO,. Since it would be interesting to examine the pH effect for catalysts
covering a wide range of molybdenum loadings, we report in this work a com-
parative study of the effect of the impregnation pH on the structure of MOO,
supported on A1203 and Si02 by a combined use of TPR, X-ray diffraction
(XRD ), and zeta potential measurements.

EXPERIMENTAL

Preparation of catalysts

Catalysts with various MOO, loadings (6-18 wt.-% ) were prepared by wet
impregnation of the supports, y-alumina (BASF D-1010, surface area= 230 m2
g-l, total pore volume at p/p0 0.9975 = 0.49 cm3 g-l, micropore volume = 0.003
cm3 g- ’) and silica (BDH, surface area = 505 m2 g-l, total pore volume at
p/p0 0.9984 =0.33 cm3 gg’, micropore volume= 0.072 cm3 g-l), both crushed
and sized to 0.15-0.25 mm, with aqueous solutions of ammonium heptamolyb-
date of appropriate concentration and at different pH (range 2-11). When
necessary, the pH was adjusted by using nitric acid or NHIOH solutions. Ex-
cess water was removed in a rotary evaporator at 60’ C under reduced pressure
for ca. 2 h. The samples were dried at 120°C for 8 h and then calcined in air,
usually at 550°C for 4 h; in some cases additional samples were calcined at
lower temperatures, 200,300 or 400°C in order to examine the genesis of the
supported molybdenum species. Calcined catalysts were kept in closed tubes.
127

Catalysts (Table 1) are denoted by the wt.-% of MOO, content of the support
(A= A&,0,; S = Si02), followed by the pH of the impregnation solution, e.g.,
Mo( 18)A-11 contains 18 wt.-% MOO, on A120, and was prepared at pH= 11.

Physicochemical techniques

Surface areas of the catalysts were determined by nitrogen adsorption (BET


method) using a Micromeritics-2100E sorptometer. X-ray diffraction patterns
were obtained with a Philips PW 1030/10 diffractometer using Cu Kcu radia-
tion with a nickel filter. Electrophoretic migration measurements were carried
out in a Zeta-Meter Instrument (Model ZM-77) using 200 mg of 2 ,um samples
dispersed in 200 ml of 10e3 A4 KC1 solution as reported in a previous study
[131.
The TPR profiles of the catalysts calcined at 500’ C were carried out in a
conventional apparatus consisting in a quartz reactor attached to a thermal
conductivity detector. The sample (250 mg) was pretreated in situ at 550°C
for 2 h, followed by cooling to room temperature, then heating up to 1000” C at
a rate of 10°C min-’ in a 70% H,/Ar gas mixture. The water produced was
trapped in 13X zeolite. Hydrogen consumption was measured with a thermal
conductivity detector. Samples were precalcined in situ at 550’ C for 2 h prior
to the TPR run. Details of the TPR measurements are given in ref. 9.

RESULTS

Surface area and XRD

Table 1 shows that the specific surface area of calcined catalysts decreased
slightly with increasing molybdenum loading, probably due to blocking of the
narrower pores of the support. A marked decrease in surface area for catalysts
prepared at high and low pH values was also observed. This effect was more
pronounced for the SiO,-supported catalysts than for the Al,O,-supported ones,
particularly at high pH. This is due to the higher microporosity of the SiOz
(see Experimental) and to its high solubility at high pH, as compared with
those of A1203 [ 141. It is well known [ 151 that at pH > 9 the smaller particles
of amorphous SiOz are readily dissolved, and subsequently are deposited on
the larger particles when oversaturation is reached. This process drastically
decreases the surface area of the SiO,. On the other hand, the moderate de-
crease in surface area of the acidic-prepared Al,O,-supported catalysts is prob-
ably caused by the solubility of A1203 at pH I3 [ 141.
The XRD patterns of all the Al,O,-supported catalysts, except the MO ( 18)A-
2, did not differ significantly from that of the A1203, which is in accordance
with the literature [ 10,16,17]. However, the diffractogram of the MO (18)A-2
catalyst calcined at 550 oC (Fig. la) showed notably the main lines of MOO,
128

TABLE 1

Properties of the supports and supported molybdena catalysts

Catalyst Impregnation SBET ZPC”


PI-I (m’gg’)

Al,O, support 230 8.85


Mo(14)A-2 2 177 5.90
Mo(14)A-5 5 160 6.05
Mo(14)A-8 8 156 5.70
Mo(l4)A-11 11 198 6.0
Mo(6)A-2 2 206 7.15
Mo(lO)A-2 2 185 5.75
Mo(18)A-2 2 158 5.30
Mo(G)A-11 11 222 7.60
Mo(lO)A-11 11 212 6.00
Mo(l8)A-11 11 188 5.20

SiO, support 505 3.0


Mo(14)S2 2 253 2.4
Mo(14)S-5 5 149 2.1
Mo(14)S8 8 136 2.5
Mo(l4)S-11 11 49 2.5
Mo(6)S-2 2 302 2.4
Mo( lO)S-2 2 242 2.4
Mo(18)S-2 2 217 2.7
Mo(G)S-11 11 158 2.5
Mo(lO)S-11 11 96 2.6
Mo(l8)S-11 11 34 2.7

“ZPC = Zero point of charge.

and Al, ( Mood),, the later phase probably being formed during the impregna-
tion step [lo]. However, attempts to confirm this genesis of Al,(MoO,), by
XRD study of MO (18)A-2 samples calcined at 300 and 400” C failed; their
diffractograms showed lines hardly distinguishable from the background of
A1203, due probably to poor crystallinity of the Al, (MOO,) 3 species.
In the case of SiOz-supported catalysts, the detection of diffraction lines
whose positions are in good agreement with crystalline MOO, occurred at rel-
atively low (6 wt.-% MOO,) loading and even at basic pH, as Figs. lb and c
show, but their relative intensities do not completely agree with Moos. These
lines were significantly more intense at acidic pH, and obviously at high mo-
lybdenum loadings. The sharpness of the lines suggests that the MOO, phase
formed is highly crystalline. In none of the SiO,-supported catalysts were lines
different from those of MOO, observed. On the other hand, the diffractograms
of the SiO,-supported catalysts only dried at 120°C were more complex than
those of catalysts calcined at 550’ C. The patterns corresponding to the dried
129

40 35 30 25 20 15 10 5
29

Fig. 1. X-ray diffraction patterns of calcined (550” C) molybdena catalysts supported on: (a)
A1203, impregnation pH = 2; (b) SiOz, impregnation pH =2; and (c) SiOB, impregnation pH = 11.
Numbers in parentheses are wt.-% MOO,.

Mo( lS)S-11 and Mo( 18)S-2 samples are shown in Fig. 2. Lines attributable
to (NH1)3HSiMo12040.5H20 (ASTM 35-242) and the isopolymolybdates
(NH1)6M07024 (ASTM 36-30) and (NH,),Mo,,O,, (ASTM 37-381) can be
appreciated, they are more intense for the acidic samples, which, effectively,
exhibited the characteristic yellow color of 12-molybdosilicic acid. This color
was still visible in the MO (18) S-2 samples calcined at 200 and 300’ C, and
changed to white-greenish after calcination at 400” C. Consistently, the XRD
patterns of such samples calcined at 200 and 300’ C were essentially similar to
that of the 120”Cdried sample, while the pattern of the Mo(18)S-2 sample
calcined at 400’ C was already similar to that exhibited for the 550” C-calcined
samples, i.e., only showing the major lines of crystalline MOO,. These results
indicate that the molybdosilicic species initially present in the dried samples
are obviously destroyed by calcination between 300 and 400°C. This is in
agreement with one previous report [18] and in disagreement with others
[ 19,201. In the diffractograms of Fig. 2 a line also appears which could be
I I I I I I I I
36 34 30 26 22 16 14 12

-28

Fig. 2. X-ray diffraction patterns of the dried (12O“C) silica-supported molybdena catalysts: (a)
Mo(l8)S-11 and (b) Mo(18)S-2. Lines assignable to (NH4)3HSiMo12040 (0); (NH,),Mo,O,,
( x ); and tridymite ( 0 ) are indicated.

_L
I I I I I
0 4 8 12 16 20

Fig. 3. Variation of the zero point of charge (ZPC) with MOO, loading for catalysts supported on
A&O, (0,. ) and SiO, (0,~). Open and full symbols correspond to catalysts prepared at pH 2
and 11, respectively.

attributed to the presence of crystalline SiOz species (tridymite and cristobal-


ite ) , although in a small proportion.

Electrophoretic measurements

The zero point of charge (ZPC) of all the calcined catalysts, obtained from
zeta potential vs. pH curves [ 131, are presented in Table 1.
For the purpose of a better comparison, Fig. 3 shows the variation of the ZPC
131

with the molybdenum loading for the catalysts prepared at pH 2 and 11. For
the alumina supported catalysts, an initial region (loadings < 10 wt.-% MOO,)
in which the ZPC decreased nearly linearly with increasing molybdenum con-
tents was observed, indicating that molybdenum is highly dispersed on A1203.
Then another region followed (loading lo-15 wt.-% MOO,) in which the ZPC
was nearly constant and close to the isoelectric point of 6.2 of bulk MOO,,
indicating that the molybdenum is forming multilayers or MOO, clusters grow-
ing vertically. Finally, for MOO, loadings 2 18 wt.-% a further small decrease
in the ZPC was appreciated, which could be attributed to the presence of
Al, ( MOO,), species (isoelectric point < 6), as detected by XRD.
For the SiO,-supported catalysts, Fig. 3 shows that the ZPC did not change
significantly with increasing molybdenum loadings, suggesting that most of
the SiO, surface is free from molybdenum, which would be present predomi-
nantly as large crystallites of MOO,.

TPR

The TPR profiles of MoO,/Al,O, and MoOJSiO, catalysts containing equal


molybdenum loadings (14 wt.-% MOO,) and prepared at different impregna-
tion pH’s are presented in Figs. 4a and b, respectively. The two sets of TPR
were clearly different.
The TPR profiles of all the AlzO,-supported catalysts (Fig. 4a) exhibited
the two reduction regions previously found [ 71. No significant changes in the
profiles with the pH were observed, except a small increase in the intensity of
the first peak at T,,,=41O”C (T,,, -temperature at which the reduction rate
is maximal) and the appearance of a small shoulder at about 480°C for the
sample prepared at pH = 2.
On the contrary, in the SiO,-supported catalysts the TPR profiles displayed
generally three reduction peaks, which were very sensitive to the impregnating
pH (Fig. 4b). The more relevant differences were: (i) the first peak,
T,,, = 490’ C, increased generally with increasing pH from 2 to 8; (ii ) the sec-
ond peak, T,,, = 6OO”C, varied oppositely to the first one; and (iii) the third
peak, T,,, = 680-7OO”C, did not change sensibly with pH, except for the ap-
pearance of a small shoulder about 750°C for the catalysts prepared at acidic
pH. For comparison, we show in the inset of Fig. 4 the TPR profile of bulk
MOO,, which exhibit two peaks with features rather similar to those of the
Mo( 14)S-2 sample, but shifted about 25 or 50°C towards higher reduction
temperatures.
The effect of molybdenum loading on A1203 at impregnation pH values 2
and 11 is shown in Fig. 5. The main differences in the TPR profiles between
samples prepared at different pH’s were observed for the 6 and 18 wt.-% MOO,
loadings catalysts. Thus the Mo(6)A-2 catalysts exhibited practically only a
reduction peak in the high temperature region (T,,, cz849” C ), while the
132

I I I I I I I I I I I I I I I I
400 600 800 1000 400 600 800 1000
T(‘C) T(=‘C1

Fig. 4. TPR profiles of 14 wt.-% MOO, catalysts supported on (a) A&O, and (b) SiO, at various
impregnation pH values. Inset: TPR profile of bulk MOO,.

Fig. 5. TPR profiles of various Mo03/A1,03 catalysts prepared at pH 2 (- - -) and 11 (---_). MOO,
contents are indicated.

MO(6)A-11 catalysts presented two well defined peaks at ca. 445 and ca. 770’C,
respectively. For the catalysts with 18 wt.-% MOO, the appearance of a new
peak ( T,,, = 495’ C ) in the preparation at pH = 2, and the absence of the peak
for the homologous preparation at pH= 11, are remarkable. The peak at
T,,, = 495 oC was also detected in the catalysts MO (14)A-2.
A similar comparison for the SiO,-supported catalysts is presented in Fig. 6.
In this case, at all the molybdenum loadings there are noticeable differences
between the TPR patterns of the catalysts prepared in basic and acidic pH.
For instance, in the TPR profile of the MO (6)S-11 only two broad peaks were
found, while the profile of the Mo(6)S2 exhibited four reduction peaks.
The quantitative results of the reduction, expressed as degree of reduction
(a = H, consumed&H, consumedth,,,, for MoO,+Mo”) of the Al,O,-sup-
ported catalysts as a function of the temperature, are given in Fig. 7. In general,
the reducibility of the Al,O,-supported catalysts increased with increasing mo-
lybdenum loading, and was higher for the one prepared at pH = 11 than for the
133

Tt’C) T(‘C)

Fig. 6. TPR profiles of various MoOJSiO, catalysts prepared at pH 2 (- - - ) and 11 (-). The
MOO, contents are indicated.

Fig. 7. Degree of reduction (LX) vs. temperature for Mo03/A1203 catalysts prepared at pH 2
(- - - ) and 11 (--). MOO, loadings are indicated.

one prepared at pH = 2. In contrast, for the SiO,-supported catalysts no sig-


nificant differences in reducibility could be appreciated among them: the ex-
tent of reduction at the end of the TPR was always total, i.e., all the molyb-
denum species present were reduced to Moo ((Y= 1).

DISCUSSION

Significant differences were observed in the TPR profiles of the A1203- and
SiO,-supported molybdena catalysts when varying the molybdenum content
and/or the pH of the impregnating solution, suggesting changes in the nature
and distribution of the molybdenum species on the surface of the catalysts.
Some of these TPR-monitored changes were, however, not detected by XRD.
For the interpretation of the TPR profiles we recall the information ob-
tained in our previous TPR study [ 71 on Al,O,-supported molybdena catalysts
at thermo-reduction conditions similar to the present ones. The first peak at
405°C was attributed to the first reduction step (Mo6+ +Mo4+ ) of octahe-
drally coordinated molybdenum species (MO,) weakly bound to A1,03, pre-
134

dominantly as multilayer molybdenum domains, with some polymolybdates in


monolayer patches. On this basis, the observed increase in the intensity of the
first peak of A&O,-supported catalysts with increasing molybdenum content
(Fig. 5)) and also with decreasing pH (Figs. 4a and 5), indicates that such
variations in the two preparational parameters lead to an increase of the pro-
portion of the octahedral reducible molybdenum species, as shown in Fig. 7.
This increase of MO, species upon decreasing the impregnation pH, can be
understood by considering that the impregnation pH controls the proportion
of MOO:- and Mo,Oz, through the equilibrium 7MoOi- +
8H+ +Mo,O& +4Hz0 and also the surface charge of the support according
to its IEP. It is well known [ 141 that the adsorption of anions is enhanced
when the support is positively charged, which is the case for Al,O, when the
pH of the impregnating solution is lower than its IEP value of ca. 8.8. Another
factor to be considered is that Al,O, preferentially adsorbs Mo,O!j’; , leaving
MOO:- in the solution into the pore as a result of a basic buffer effect [ 111.
Thus, at pH I5 and high concentration of molybdenum in the impregnating
solution, although the adsorption of the predominant Mo,O;; species [6] is
favoured, a considerable amount of molybdenum remains in the solution since
its concentration is relatively high. Consequently, when the solvent is removed
the unadsorbed Mo,O!& species will precipitate and/or polymerize, and upon
drying and calcination they will form a heterogeneous deposit of polymeric
molybdenum species on the A1203, and even MOO, crystallites, as occurs in the
case of the MO (14)A-2 and MO (18)A-2 catalysts. The presence of crystalline
MOO,, together with A12(Mo04)3, in the MO (18)A-2 was confirmed by the
XRD results (Fig. 1) , and by the appearance in the TPR profile of a new peak,
T,,, = 490’ C, well defined and relatively intense. This peak was also observed
in similar molybdenum catalysts containing phosphorus [9,10]. In the
MO (14)A-2 catalyst the presence of crystalline MOO, was not detected by XRD,
likely due to the small size of the aggregates or crystals, but it was revealed by
TPR by the small shoulder exhibited at ca. 480 oC (Fig. 4a).
In the preparations at pH 2 8 the solutions contain predominantly MOO:-
species. At these pH values the alumina is negatively charged and therefore no
adsorption of anions occurs. Thus the unadsorbed species deposited on the
alumina after drying will be essentially MOO!-. This then leads to a minor
proportion of molybdenum in polymeric form or as MoOB crystallites, as sug-
gest the TPR profiles of the Mo( 14)A-11 and Mo( 18)A-11 catalysts (absence
of a peak at ca. 480-490°C)) and hence to less reducibility, than for the ho-
mologous catalysts prepared at pH < 8 (Fig. 7). This is in line with the findings
derived from the ZPC results (Fig. 3), that the molybdenum dispersion on
alumina-supported catalysts was generally lower at pH 11 than on those pre-
pared at pH 2.
Comparing the TPR profiles of the SiO,-supported to the Al,O,-supported
catalysts, the absence of the peak at T,,, = 410’ C in the silica catalysts indi-
135

cates that, in contrast to the alumina catalysts, they do not contain Mo, species
in monolayer and/or multilayer form. The explanation is that on SiO,
(IEP z 2.5) practically no adsorption of anions occurs since its surface is neg-
atively charged in almost the whole pH range studied. In addition, the inter-
action of the oxomolybdenum species with SiOZ is very weak.
In contrast with the AlaOs-supported catalysts, all the SiO,-supported cat-
alysts showed the peak at T,,, - 490’ C, which is indicative of a discrete MOO,
phase. The similarity of the TPR of bulk MOO, (peaks at 625 and 760” C) with
the TPR profiles of some SiO,-supported catalysts, e.g. the MO (14) S-2 cata-
lysts (peaks at ca. 600 and ca. 700’ C ) , can clearly be seen by comparing Fig. 6
and the inset of Fig. 4b. The presence of small crystals of MOO, in the silica
catalysts was corroborated by the XRD results (Fig. lb and c). This MOO,
phase, which gave rise to the peak at ca. 490” C, developed further with increas-
ing pH of impregnation (Fig. 4b). However, in the alumina supported catalysts
the effect of the pH was contrary, as Fig. 4a shows. Another significant differ-
ence between the A1203- and SiO,-supported catalysts is that for the latter
there appeared a second peak ( T_ - 600 oC ), very close to that corresponding
to the first reduction step of bulk Mo03. This peak at ca. 600°C intensified
with decreasing pH of impregnation, conversely to the trend followed by the
peak at ca. 490°C (Fig. 4b). We conclude that both peaks result from the ex-
istence of two different MOO, bulk phases which may differ in the size of crys-
tal or aggregate, and whose proportions are strongly dependent on the impreg-
nation pH. The MOO, phase predominant at acidic pH, which gives rise to the
peak at 6OO”C, may have a larger crystal size than that predominant at basic
pH values. This is supported by the fact that the diffraction patterns of the
catalysts prepared at pH 2 exhibited a higher sharpness than those prepared
at pH 11, as Fig. 1 shows. This Moo3 phase which forms predominantly at
acidic pH values may result from the unadsorbed Mo,O& species, and also
from the decomposition of the molybdosilicic acid which was formed during
the impregnation. However, the more abundant MOO, phase at basic pH
( T,,, = 490’ C ), apparently more disperse than that predominant at acidic pH,
may result from unadsorbed MOO;- anions, which are predominant at basic
pH. The lower temperature of reduction of the MOO, phase at basic pH is
thought to be caused by the relatively smaller size of its crystals, as compared
to that in the catalysts prepared at basic pH. An additional reason could be
that the MOO, phase may be finely mixed with the precipitated SiOa proceed-
ing from the partial solubilization occurring during the impregnation at pH
> 9.
Finally, the peak at 680-700 oC in the TPR profiles is ascribed to the second
step of reduction of the MOO, species in the case of SiO,-supported catalysts,
and to the partial reduction of molybdenum species strongly bound to the sup-
port in A&O,-supported catalysts.
136

ACKNOWLEDGEMENTS

The financial support of the DGICYT (Spain) Project PB87-0261,


FONDECYT Project 0761-89 (Chile), and the Program of Scientific Cooper-
ation between the CSIC (Spain) and the CIQ (Cuba is gratefully acknowledged.

REFERENCES

1 Y. Okamoto and T. Imanaka, J. Phys. Chem., 92 (1988) 7102.


2 L. Wang and W.K.J. Hall, J. Catal., 66 (1980) 251.
3 S. Kasztelan, J. Grimblot, J.P. Bonnelle, E. Payen, H. Toulhoat andY. Jacquin, Appl. Catal.,
7 (1983) 91.
H. Jeziorowski and H. Knijzinger, J. Phys. Chem., 83 (1979) 1166.
C.P. Cheng and G.L.J. Schrader, J. Catal., 60 (1979) 276.
N.P. Luthra and W.C. Cheng, J. Catal., 107 (1987) 154.
R. Lopez Cordero, J. Lbzaro, J.L.G. Fierro and A. Lopez Agudo, in F. Cossio, 0. Bermudez,
G. de1Angel and R. Gomez, (Editors), Actas XI Simposio Iberoamericano de Catilisis, IMP-
SIC, Junio 1988, Mexico D.F., pp. 563-570; and references therein.
8 P. Arnoldy and J.A. Moulijn, J. Catal., 93 (1985) 38.
9 R. Lopez Cordero, N. Esquivel, J. Lbzaro, J.L.G. Fierro and A. Lopez Agudo, Appl. Catal.,
48 (1989) 341.
10 R. Lopez Cordero, S. Lopez Guerra, J.L.G. Fierro and A. Lopez Agudo, J. Catal., 126 (1990)
8.
11 P. Sarrazin, B. Mouchel and S. Kasztelan, J. Phys. Chem., 93 (1989) 904.
12 H.M. Ismail, C.R. Theocharis and M.I. Zaki, J. Chem. Sot., Faraday Trans. 1, 83 (1987)
2835.
13 F.J. Gil Llambias and A.M. Escudey Castro, J. Chem. Sot. Chem. Commun., (1982) 478.
14 J.P. Brunelle, Pure Appl. Chem., 50 (1986) 329.
15 R.K. Iler, The Chemistry of Silica, Wiley, New York, 1979, p. 40.
16 H.M. Ismail, C.R. Theocharis, D.N. Waters, M.I. Zaki and R.B. Fahim, J. Chem. Sot., Far-
aday Trans. 1,83 (1987) 1601.
17 J. Valyon, M. Henker and K.P. Wendlandt, React. Kinet. Catal. Lett., 38 (1989) 265.
18 C. Rochiccioli-Deltcheff, M. Amirouche, M. Che, J.M. Tatibouet and M. Fournier, J. Catal.,
125 (1990) 292.
19 L. Rodrigo, K. Marcinkowska, A. Adnot, P.C. Roberge, S. Kaliaguine, J.M. Stencel, L.E.
Makovsky and J.R. Diehl, J. Phys. Chem., 90 (1986) 2690.
20 S. Kasztelan, E. Payen and J.B. Moffat, J. Catal., 112 (1988) 320.

You might also like