Sei sulla pagina 1di 5

Bismuth Molybdate Catalysts

Kinetics and Mechanism of Propylene Oxidation

George W. Keulks', Mich I P. R k, and Chelliah Daniel


Department of Chemistry and Laboratory f o r Surface Studies, Uniiiersity of
Wisconsin-Milwaukee,Milwaukee, W i s . 53201

Oxidation of propylene to acrolein over bismuth molybdate catalysts was studied in


a flow reactor a t atmospheric pressure and a t temperatures between 400" and 460C.
The kinetics were determined as first order in propylene and zero order in oxygen.
Over this temperature range, the activation energy was 29 kcal per mole. The side
products: acetaldehyde, formaldehyde, ethylene, carbon monoxide, and carbon dioxide,
were formed almost exclusively in a consecutive pathway via acrolein or its surface
species precursor when excess oxygen was present, A parallel pathway from propylene
for the formation of carbon dioxide became important when there was an oxygen
deficiency. A gas-phase, homogeneous oxidation of acrolein also became important
when the postcatalytic volume of the reactor was increased.

T h e catalytic oxidation of olefins, using bismuth pathway (Sampson and Shooter, 1965). The work of
molybdate catalysts, has received considerable attention Gorshkov et al. (1968), however, indicates that, in addition
during the past decade. Studies involving olefins were to acrolein, other intermediates, which do not have acrolein
principally concerned with the oxidation of propylene to as a precursor, produce carbon dioxide when oxidized over
form acrolein and with the oxidative dehydrogenation of bismuth molybdate. Thus, carbon dioxide may be pro-
n-butenes to form 1,3-butadiene. duced by a number of pathways as well as from a number
The mechanism by which propylene reacts, a t least of intermediates.
initially, is firmly established and extensively discussed In view of the uncertainty regarding the origin of the
in recent reviews by Sachtler (1970), Sampson and Shooter various side products of the reaction, the present study
(1965), and Voge and Adams (1967). The rate-determining was undertaken with the particular objective of gaining
step in the oxidation is the abstraction of an allylic hydro- further insight into the mechanism of the reaction. In
gen atom by surface oxygen t o form a symmetric allyl addition, the authors decided to examine the kinetics of
intermediate. A subsequent abstraction of a second hydro- the catalytic reaction by carefully analyzing the product
gen atom from either terminal carbon atom of the allyl distribution and by paying particular attention to min-
intermediate followed by oxygen atom incorporation yields imizing competitive, homogeneous, gas-phase reactions.
acrolein. Keulks (1970) used 1802 to show that essentially
Experimental
all of the oxygen atoms of the bismuth molybdate catalyst
participate in the oxidation and serve as the source of Reactor. The kinetic experiments were run in a single-
oxygen for the incorporation into the products, confirming pass, integral-flow reactor a t 350" to 475C and atmo-
the suggestions of Batist et al. (1967) and Callahan et spheric pressure. The reactors consisted of a length of
al. (1970). 10-mm borosilicate glass tubing. I t was heated by a tubular
The oxidation of propylene over bismuth molybdate furnace, whose temperature was controlled to within 10.5'
yields. in addition to acrolein, considerable amounts of by a proportional temperature controller. The catalyst
carbon monoxide, carbon dioxide, and water, as well as was retained by borosilicate glass-wool plugs and the fur-
lesser amounts of ethylene, formaldehyde, and acet- nace was adjusted so that the catalyst section was in
aldehyde. Kinetic studies show the overall reaction to the center. T o minimize the possibility of a homogeneous
be first order in propylene pressure and zero order in reaction in the postcatalytic zone, the reactor was con-
oxygen pressure (Adams et al., 1964; Gelbshtein et al., structed out of 2-mm capillary tubing below the portion
1965; Gorshkov e t al., 1970) which is consistent with containing the catalyst.
the proposed mechanism. Oxygen (Airco, 99.6%), propylene (Baker, 95%), helium
A topic of chief concern in several investigations is diluent (Airco, 99.99%), and ethane (Baker, 99%), which
the origin of the carbon dioxide formed during the reaction. was used as an internal standard, were further purified
I t appears now that carbon dioxide is formed from both by passage through molecular sieve traps and micron
propylene and acrolein in a parallel-consecutive type of filters. The flow rates of the four gases were monitored
by flow transducers (Gow-Mac Co.), and were controlled
' To whom correspondence should be addressed. with Hoke microneedle valves. The gas mixture was sub-

138 Ind. Eng. Chem. Prod. Res. Develop., Vol. 10, No. 2, 1971
was used in the kinetic experiments. Before any kinetic
data were obtained, the catalyst was lined-out for 18
hr a t 425'C with a flow of propylene, oxygen, and helium
in the ratio 40:40:80 cm' (STP) per min.
Reaction Kinetics

The kinetic order with respect to oxygen was determined


by varying the oxygen partial pressure while maintaining
the propylene partial pressure at 0.25 atm, and the total
flow at 160 cm' (STP) per min. As shown in Figure
1, the oxidation of propylene is independent of the oxygen
partial pressure over a range of 0.15 to 0.8 atm. At partial
pressures of oxygen below 0.15 atm, the catalyst rapidly
lost activity. This is in agreement with the reported results
of Callahan et al. (1970), Dalin et al. (1962), and Gelbsh-
tein et al. (1965) who observed a similar loss in activity
at low oxygen partial pressures. This rapid loss in activity
in a reducing atmosphere is owing to the reduction of
Mo6- to Mo4 which results in the depletion of active
lattice oxygen (Callahan et al., 1970).
Because the catalyst rapidly lost activity in the presence
of excess propylene, the kinetic order with respect to
I I 1 1 I I I I I
propylene was determined in excess oxygen. This was
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
accomplished by varying the contact time while main-
P ,atm taining the propylene partial pressure at 0.25 atm and
02 the oxygen : propylene : helium ratio at 1.5:1:1.5.The
Figure 1. Percentage conversion of propylene vs. oxygen first-order rate plots of -In (1 - x ) , where x is the fraction
partial pressure of propylene converted vs. reciprocal space velocity, are
Temp, 425C; pressure (CIH~), 0.25 atm; total flow, 160 cm' (STP) per shown in Figure 2. As indicated, the reaction conformed
min; catalyst wt, 3 grams to first-order kinetics over the temperature range 400"
to 460" C. The apparent activation energy for the reaction
was computed as 29 f 5 kcal per mole from the Arrhenius
sequently homogenized prior to entering the reactor. equation using the integral reaction rates determined from
Ethane was used as an internal standard because of its the first-order rate plots. Considering the complexity of
demonstrated inertness under reaction conditions and its the reaction along with the experimental error involved
convenient chromatographic retention time. I t was used in such measurements, this value is in reasonable agree-
in conjunction with known thermal response factors (Dietz, ment with those reported by Adams et al. (1964) and
1967) to obtain absolute calibrations of the gas chro- more recently by Peacock et al. (1969).
matograph for each component of the reaction effluent in
j terms of peak areas vs. micromoles.
Gas chromatographic analysis of the reactor exit gas
was accomplished by diverting the effluent through a 0.6-
10 -
cm3 loop of a gas sampling valve (Carle Instruments Co.,
No. 2014). The gas chromatograph contained two columns: 9 -
silanized Porapak Q (Waters Assoc.) which provided the
separation for C O P ,methane, ethylene, ethane, propylene, a -
water, formaldehyde, acetaldehyde, and acrolein; and
molecular sieve 5A, which provided the separation of CO 7 -
and 0'.
The propylene conversion was calculated using the
internal standard according to Hall et al. (1960), and
by using the sampling valve to analyze the feed mixture
prior to its entrance into the reactor, and then comparing
the peak areas of propylene before and after reaction.
The consistent agreement of these two methods in cal- I / f
culating the various conversions was indicative of the
lack of an appreciable volume change during the reaction.
Catalyst Preparation and Pretreatment. The bismuth
molybdate catalyst used in the experiments was prepared
by coprecipitation from solutions of bismuth nitrate and 0.2 0.4 0.6 0.8 1.0 1.2 14 1.6 1.8 2.0
ammonium molybdate, according to the method of Adams
et al. (1964). I t had a surface area of 2.5 meters' per A/Fi rnin-m2/rnoles x lo2
gram, as determined by the B E T method using nitrogen Figure 2. First-order rate plots
at -195" C as the adsorbate. 0 460"C A 425" C
Approximately three grams of 40- to 100-mesh particles 450"C + 400C

Ind. Eng. Chem. Prod. Res. Develop., Vol. 10, No. 2, 1971 139
products are allowed to remain a t elevated temperatures
Table I. Effect of Oxygen Partial Pressure
in the postcatalytic zone, they can undergo homogeneous
on the Product Distribution
reactions.
[Temp, 425C; pressure (CiH6),0.25 atm; total flow; 160 In an effort to gain further insight into the thermal
em' (STP) per min; catalyst wt, 3 grams] stability of the reactants and some of the products, a
Moles per 100 moles C I H ~converted reactor was constructed of 10-mm borosilicate glass tubing.
Pressure
I t contained no catalyst and was packed with 4-mm
(0.).atm YO CO CO, C,H, C~HIO CiHiO
borosilicate glass beads. As expected, under conditions
0.125 19 13 23 1 D 73 similar to those used for catalytic studies, propylene (40
0.250 22 19 24 2 7 80
0.312 20 20 22 2 7 76 cm' per rnin in 160 cm' per min helium) and propylene
0.375 21 17 25 2 7 rr
ID : oxygen : helium mixtures (40:60:60 cm' per min)
0.500 20 18 24 2 7 nr
ii underwent no homogeneous reaction a t 425" C. Similarly,
0.625 23 25 23 2 6 73 acrolein (40 cm3 per rnin in 180 cm3 per min helium),
0.750 24 26 26 2 7 73 in the absence of oxygen, was inert under these conditions.
Mixtures of oxygen and acrolein, however, exhibited
Table I I . Pulse Reactor Studies of Propylene extensive reaction a t 425" C, in some cases with explosive
Oxidation Components force. A qualitative study of this process was made by
Pulse components Temp, C Observed products passing a 1:l mixture of oxygen and acrolein through
the reactor a t varying contact time. As expected, the
C Hi, 300-500 None
CIHi + 01 300-500 None extent of conversion of acrolein decreased with decreasing
C ,HI0 300-485 None contact time. More significantly, however, although the
GH,O + 02 300-500 CO, Cor, CPH,,H>O,CHqCHO relative amounts of carbon dioxide, ethylene, and water
CHiCHO 200-500 CO + CH, (traces) also decreased with increasing flow rate, the amounts of
CH iCHO + 0, 200-500 CO, COP, HCHO, H20, CH,
formaldehyde and acetaldehyde exhibited corresponding
increases with decreasing contact time, suggesting their
roles as intermediates in the overall homogeneous decom-
position of acrolein (Kusuhara, 1961).
Product Distributions
The thermal stability of the various products observed
The product distribution, in terms of moles per 100 in the oxidation of propylene in a flow reactor also was
moles of propylene converted, for various oxygen partial studied in a pulse reactor. The helium flow was adjusted
pressures is given in Table I. These results show that to give contact times in the hot zone comparable to those
the product distribution does not vary significantly with in the flow reactor. The results of these studies are given
the oxygen partial pressure. The amount of acrolein formed in Table 11.
as a function of temperature and contact time is shown The results obtained with a reactor having a large
in Figure 3. The corresponding changes in the product postcatalytic zone, the results obtained with a flow reactor
distribution are shown in Figure 4. The yield of acrolein containing glass beads, and the pulse reactor studies clearly
decreases with both an increase in temperature and an indicate that homogeneous reactions between oxygen and
increase in contact time while an increase is observed
in the yield of the various products.
These results suggest that the various products are pro-
duced from an intermediate in a reaction step that kineti-
cally does not depend upon the gaseous partial pressure
of oxygen. Thus, as indicated earlier, the reactive oxygen
in this reaction step probably comes from the bismuth
molybdate catalyst itself. I n addition, these results suggest
that acrolein, or some surface species which is a precursor
t o acrolein, is the intermediate responsible for the forma- z_
Lu
tion of the various products observed in the oxidation.
Homogeneous Reactions and Thermal
Stability of the Products

The product mixtures resulting from the oxidation of


propylene conceivably may undergo a variety of homo-
geneous reactions in the postcatalytic zone of the reactor
(McCain and Godin, 1964). T o examine the importance
of such homogeneous reactions in more detail, a number
of experiments were conducted with a reactor in which
the volume of the postcatalytic zone was increased by I I I I 1 I 1 1 1
replacing the capillary tubing with 10-mm tubing. The 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 :0
amount of acrolein decreased and also decreased sig-
nificantly with an increase in the oxygen partial pressure A/Fi min-m2/rnoles x102
as shown in Figure 5 . Also, the decrease in the amount Figure 3. Effect of contact time and temperature on yield
of acrolein observed with increasing partial pressure of of acrolein
oxygen in the feed was accompanied by concomitant in- W 460" C 0 425" C
creases of all of the various side products. Thus, if the A 450 C X 400C

140 Ind. Eng. Chem. Prod. Res. Develop., Vol. 10, No. 2, 1971
100 1 their relative amounts were essentially identical t o those
observed for the oxidation of propylene. These results
are in agreement with those reported by Gorshkov et
90
a0
tL i9
4 8
al. (1968), except that in the case of acrolein they observed
acrylic acid and did not observe acetaldehyde. Their study
of a number of other side products observed in the oxida-
tion of propylene also showed that the oxidation of these
side products occurs by consecutive-parallel pathways.
These authors conclude that the main sources of CO?
formation are acids and that the main sources of CO
formation are aldehydes.
The results shown in Figures 3 and 5 , together with
those of Gorshkov et al. (1968), and those from the acrolein
oxidation experiments offer strong evidence for the conclu-
sion that all of the side products observed during the
- 1 oxidation of propylene, including virtually all of the CO
I l l 1 1 1 1 1 1 1
and CO,, arise from subsequent oxidation of acrolein or
0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 20 its surface species precursor. However, this appears . t o
be true only when propylene is reacted with excess oxygen.
A/Fi, min-mymoles x 1 0 2 In the absence of excess oxygen, a pathway for CO,
Figure 4. Variation in the product distribution at 425C formation directly from propylene apparently becomes
as a function of the space velocity available. This is suggested by the observation that when
0 Acetaldehyde A CO the oxygen partial pressure is greatly reduced, the forma-
Ethylene + CO? tion of COYpredominates. Unfortunately, the rapid loss
of catalytic activity in the highly reducing atmosphere
the various side products can occur and, indeed, contribute prevented us from obtaining any quantitative results.
significantly to the observed product distribution if proper However, this observation is in agreement with the results
precautions are not taken to minimize these reactions. of Peacock et al. (1969). They report that in the absence
The decrease in the amount of acrolein observed with of oxygen, the reaction of propylene with bismuth
increasing oxygen partial pressure (Figure 5 ) and the con- molybdate produces CO, predominately. They have also
comitant increases for all of the various side products shown that acrolein, on the other hand, in the absence
indicate that acrolein may serve as the precursor for these of oxygen produces twice as much CO as CO? and in
products in the homogeneous reaction.
Catalytic Reactions

The results presented in Table I , in which a reactor


ao c
l
T~ 100
k
having a minimal postcatalytic volume was used, are
believed to be indicative of the product distribution arising
mainly from the catalytic reaction. As shown in Figures
3 and 5, the curves for acrolein extrapolate to 100 moles
70 c \
of acrolein formed per 100 moles of propylene reacted
a t infinite space velocity. Thus it appears that the primary
catalytic process is

CH,=CHCH, + 2 O(cata1yst) -+

CH,=CHCHO + H20
As noted earlier, the bismuth molybdate catalyst serves
as the source of oxygen for the reaction.
T o gain further insight regarding the formation of the
various side products, a series of experiments was per-
formed in an effort t o determine the relative reactivity
of some of the observed products under our reaction condi-

2oiY
tions. When a 1.5:1.0:3.0 mixture of oxygen : ethylene
: helium (total flow = 180 cm' per min) was passed over
the bismuth molybdate catalyst a t 425" C, no conversion
to any products was observed. Similar experiments with
an oxygen : carbon monoxide mixture indicated that the
rate of oxidation of carbon monoxide over the catalyst
' OOOL ' 0.1 0.2 0.3 0.4 0.5
is also negligibly slow a t 425C. When a 2:1:3 mixture
Partiol Pressure of Oxygen
of oxygen : acrolein : helium (total flow approximately
180 cm' per min) was passed over the catalyst a t 425"C, Figure 5. Effect of oxygen partial pressure on acrolein yield.
however, 85% conversion of acrolein was observed. The Large postcatalytic volume in reactor
oxidation products (carbon monoxide, carbon dioxide, Temp, 425C; pressure (CyHs), 0.167 atm; total flow, 240 cm (STP)
water, ethylene, formaldehyde, and acetaldehyde) and per min; catalyst wt, 3 grams

Ind. Eng. Chern. Prod. Res. Develop., Vol. 10, No.2, 1971 141
the presence of excess oxygen only nearly equal amounts Dalin, M. A., Lobkina, V. V., Abaev, G. N., Serebryakov,
of CO and CO?. Thus, the enhanced COz production under B. R., Plaksunova. S. L., Dokl. Akad. Nauk SSSR,
reducing conditions cannot be explained by an enhanced 145, 1058 (1962).
consecutive reaction of acrolein. Instead, it is suggested Dietz, W. A., J . Gas Chromatogr., 5, 68 (1967).
that under these conditions a parallel pathway for com- Gelbshtein, A. I., Bakshi, Yu. M., Stroeva, S. S., Kulkova,
plete combustion directly from propylene becomes N. V., Lapidus, V. L., Sadovskii, A. S.,Kinet. Katal.,
6, 1025 (1965).
important. Gorshkov, A. P., Gagarin, S.G., Kolchin, I. K., Margolis,
I n conclusion, it may be stated that the primary L. Ya., Neftekhimiya, 10, 59 (1970).
catalytic reaction of propylene oxidation over bismuth Gorshkov, A. P., Kolchin, I. K., Gribov. A. M., Margolis,
molybdate is the formation of acrolein. The side products L. Ya., Kinet. Katal., 9, 1086 (1968).
formed result from a subsequent oxidation of the acrolein Hall, W. K., MacIver, D. S., Weber, H. P., Ind. Eng.
or its surface species precursor. The selectivity for acrolein Chem., 52, 421 (1960).
is also complicated by the reactor design and the propylene Keulks, G. W., J . Catal., 19, 232 (1970).
: oxygen ratio: a large postcatalytic volume increases Kusuhara, S.,Reu. Phys. Chem. (Japan), 31, 34 (1961).
significantly the occurrence of homogeneous reactions; and McCain, C. C., Godin, G. W., Nature, 202, 692 (1964).
Peacock, J. M., Parker, A. J., Ashmore, P. G., Hockey,
a deficiency of oxygen enhances the parallel pathway from
J. A., J . Catal., 15, 398 (1969).
propylene for complete combustion to COz. These factors Sachtler, W. M. H., Catal. Reo., 4, 27 (1970).
need t o be carefully considered in obtaining the maximum Sampson, R. J., Shooter, D., in Oxidation and Combus-
selectivity for acrolein formation over bismuth molybdate tion Reviews, C. F. H. Tipper, Ed., Vol. 1, Elsevier,
catalysts. Amsterdam, The Netherlands, 1965, pp 223,256.
Voge, H. H., Adams, C. R., Aduan. Catal. Relat. Subj.,
Literature Cited
17,151 (1967).
Adams, C. R., Voge, H. H., Morgan, C. Z., Armstrong,
W. E., J . Catal., 3, 379 (1964). RECEIVED
for review July 23, 1970
Batist, Ph. A., Kapteijns, C. J., Lippens, B. C., Schuit, ACCEPTED January 28, 1971
G. C. A,, ibid., 7, 33 (1967). Partial financial assistance for this work was provided by a National
Callahan, J. L., Grasselli, R . K., Milberger, E. C., Strecker, Science Foundation-Departmental Science Development Program
H. A., Ind. Eng. Chem. Prod. Res. Develop., 9, 134 Grant. Presented at the Division of Physical Chemistry, 159th
(1970). Meeting, ACS, Toronto, Canada, May 1970.

Hydroformylation of 1 -Pentene
and 1-Hexene in Presence of
Phosphine-Modified Ca ta Iys t s

Wolfgang Rupilius, John J. McCoy, and Milton Orchin


Department of Chemistry, University of Cincinnati, Cincinnati, Ohio 45221

I t is now well-known that the addition of trialkylphos- ( P R J ? , are present, or, when excess butyl phosphine is
phines to the conventional hydroformylation catalyst can present with small concentrations of cobalt. I n fact, the
lead to as much as 88 to 90% of straight alcohol (Slaugh controlling conditions determining the selectivity to
and Mullineaux, 1966, 1968; Tucci, 1968) compared to straight-chain isomer are probably the positions of the
the 60 to 70% in the absence of phosphines. I t is also equilibria (Bianchi et al., 1969; Piacenti et al., 1970; Szabo
well-established that the presence of phosphines retards et al., 1968)
the rate of hydroformylation. The early claim (Tucci, )Z 2 C O ~ ( C O ) ~+
C O ~ ( C O ) ~ ( P R+~CO PRPR,
~ (1)
1968) that the hydroformylation of terminal olefins in
the presence of n-BusP leads to high proportions of HCo(C0)3PR, + CO 2 HCo(C0)d + PR, (2)
straight-chain isomers independent of temperature (150 In the presence of an excess of P R I , the active catalyst
to 180C ) , catalyst concentration, or carbon monoxide under most conditions is H C O ( C O ) ~ P R +This catalyst
partial pressure now appears to be valid only when large is responsible for high selectivity to straight-chain product.
concentrations of the bisphosphine complex, C O ? ( C O ) ~ - Under conditions where H C O ( C O )may~ be present, consid-
erably more branched-chain product is formed. This work
To whom correspondence should be addressed. shows that the relative amounts of the different catalysts

142 Ind. Eng. Chem. Prod. Res. Develop., Vol. 10, No. 2, 1971

Potrebbero piacerti anche