Vous êtes sur la page 1sur 15

Journal of Molecular Catalysis A: Chemical 134 Ž1998.

47–61

Review
Shape-selective hydrogenation and hydrogen transfer reactions
over zeolite catalysts 1
E.J. Creyghton ) , R.S. Downing
Laboratory of Organic Chemistry and Catalysis, Delft UniÕersity of Technology, Julianalaan 136, 2628 BL Delft, Netherlands
Received 11 April 1997; accepted 25 November 1997

Abstract

A review of the application of zeolite systems in shape-selective hydrogenation and hydrogen transfer reactions is
presented. Two different types of catalytic systems are discussed. The first consists of metal clusters or coordination
complexes encapsulated in the micropores of a zeolite employing hydrogen as the reductant. Recent developments, such as
the application of zeolite-containing composite catalysts are included. In the second system, zeolites are applied as catalyst
in hydrogen-transfer reactions using a secondary alcohol as hydrogen donor in the Meerwein–Ponndorf–Verley reduction.
q 1998 Elsevier Science B.V. All rights reserved.

Keywords: Zeolite; Shape-selectivity; Hydrogenation; Hydrogen transfer; Meerwein–Ponndorf–Verley reduction; Metal cluster; Metal
complex

1. Introduction molecules which are small enough to enter the


zeolite pores. Because of their adsorption prop-
Employing a zeolite as carrier for a noble erties, zeolites with a hydrogenation function
metal enables the shape-selective properties of will not be the catalysts of choice when selectiv-
the microporous structure to be conferred on the ity to an intermediate product is required. How-
hydrogenation function. In this way, highly se- ever, for other types of selective hydrogenation,
lective hydrogenation catalysts can be prepared, the combination noble metalrzeolite is expected
many examples of which have been described. to be beneficial.
As far as we know, however, none of these In competitive hydrogenation reactions, noble
systems are operated on an industrial scale. metal containing zeolites are recognised as use-
The use of metal-loaded zeolites in shape- ful catalysts. As a result of reactant shape selec-
selective hydrogenations is of course limited to tivity, only those molecules which are small
enough to reach the catalytic site are expected
to be converted. Furthermore, when the sizes of
)
Corresponding author. Present address: Shell Research and the competing molecules are in the same range
Technology Center Amsterdam, Badhuisweg 3, 1031 CM Amster- as the pore-size of the zeolite, mass-transport
dam, Netherlands.
1
Dedicated to Professor Herman van Bekkum on the occasion selectivity can occur which is a result of differ-
of his 65th birthday. ences in the diffusivities of the competing reac-

1381-1169r98r$19.00 q 1998 Elsevier Science B.V. All rights reserved.


PII: S 1 3 8 1 - 1 1 6 9 Ž 9 8 . 0 0 0 2 2 - 3
48 E.J. Creyghton, R.S. Downingr Journal of Molecular Catalysis A: Chemical 134 (1998) 47–61

tants. It should be kept in mind, however, that of carbonyl compounds. The activity of these
the competing molecules can influence each catalysts is related to their Lewis acid andror
other’s transport. basic properties. Some remarkable examples of
In parallel hydrogenation reactions, chemo-, shape-selective conversions resulting in high
regio- and stereoselectivity may all be achieved stereoselectivities have recently been reported.
with zeolite-based catalysts. Reactants having
molecular dimensions close to those of the zeo-
lite pores are expected to approach the metal 2. Shape-selective hydrogenation of olefins
cluster ‘end-on’. Thus, in the case of a , b-un- and arenes over zeolite-encapsulated metal
saturated aldehydes, this is expected to result in clusters
chemoselective hydrogenation to the unsatu-
rated alcohol. Similarly, terminal olefins are 2.1. Gas phase reactions
expected to be converted more easily than inter-
nal double bonds Žregioselectivity.. The encapsulation of noble metal clusters in
Stereoselectivity may occur as a result of the the micropores of a zeolite combines the possi-
confined space in which the reactions are con- bility of shape selectivity with a hydrogenation
strained to proceed. In the hydrogenation of function. This was first demonstrated in 1960
substituted cyclohexanones, for example, the by Weisz and Frilette w10x and Frilette et al. w11x
side from which hydrogen is being transferred of the Mobil Oil. A shape-selective catalyst was
Žaxial or equatorial attack. determines whether made by the incorporation of platinum in zeolite
the cis- or trans-alcohol is formed w1x. When LTA. In a mixture of equal volumes of 1-butene,
the space around the reactant is limited, adsorp- isobutene and hydrogen, at 258C with a contact
tion of the carbonyl on the metal cluster will time of ca. 0.1 s, only the 1-butene was con-
occur from the most favourable side, thereby verted to butane, while approximately equal
minimizing steric repulsion. Thus, because of quantities of both olefins were converted over a
steric constraints, stereoselectivity is expected. platinum on alumina reference catalyst. Full
In fact, this example is closely related to re- experimental details on both the catalyst prepa-
stricted transition-state selectivity. ration and the catalytic reaction were published
Some zeolites, such as ETS-10 w2x and beta in 1962 w12x. As none of the available cationic
ŽBEA. w3–6x, can theoretically be synthesised in or anionic platinum species can enter the 4.5 A ˚
homochiral forms; however, this has not yet cage windows, an aqueous solution of
been achieved w7x. Encapsulation of metal clus- platinumŽII. tetra-amine chloride was added di-
ters in the homochiral forms could yield a cata- rectly to the zeolite synthesis mixture and co-
lyst which is potentially capable of asymmetric crystallised. Removal of extra crystalline plat-
hydrogenation. An alternative technique, the inum complex by repeated ion-exchange with
modification of the interior of a zeolite with a aqueous calcium chloride was essential for the
chiral reagent can induce a chiral environment absence of isobutene hydrogenation activity.
with potential for enantioselective hydrogena- Diffusion of molecules through the pores of a
tions, as recently reported w8,9x. zeolite takes place in a regime which is called
This article presents an overview of shape- configurational diffusion w13x. It deviates from
selective hydrogenations using metal-loaded ze- classical molecular and Knudsen diffusion be-
olites. The examples illustrate both the potential cause the sizes of the molecules are in the same
and limitations of these catalytic systems. The range as the pore diameter of the zeolite chan-
review includes the recent application of zeo- nels. Moreover, very small changes in the rela-
lites as new recyclable solid catalysts for the tionship between pore diameter and molecular
Meerwein–Ponndorf–Verley ŽMPV. reduction dimensions can induce large differences in the
E.J. Creyghton, R.S. Downingr Journal of Molecular Catalysis A: Chemical 134 (1998) 47–61 49

diffusion coefficients of the reactants involved. reactants and products involved. As soon as
These large differences belong to the fundamen- these equilibrium conditions are reached, even
tal effects that give rise to shape selectivity, as though the interior of the zeolite is not saturated
shown by Chen and Weisz w14x. The diffusion with propane, additional propene entering the
coefficient of trans-2-butene in A-type zeolite catalyst will exit unconverted.
is about 200 times greater than that of cis-2- Steinbach and Minchev w16x examined the
butene, even though these two molecules vary shape-selective properties of NirCa-LTA by the
˚ Hydrogena-
in their critical size by only 0.2 A. single component hydrogenation of 1-hexene
tion of a mixture of cis- and trans-2-butene, and of 2,3-dimethyl-2-butene at 358C. No shape
over the same PtrLTA catalyst as discussed selectivity was observed for catalysts reduced in
above, resulted in a much higher hydrogenation flowing hydrogen. This was caused by metal
rate for the trans-isomer, indicating that the migration from the interior of the zeolite to the
selectivity is controlled by configurational diffu- external surface during the reduction process,
sion. Based on first-order kinetics, the relative even at moderate reduction temperatures. Mi-
rate constant of the trans-isomer was calculated gration of nickel could be avoided, however, by
to be 3.3 to 7 times higher as compared to that using sodium in liquid ammonia as the reducing
of the cis-isomer. Moreover, on non-micropor- agent of the Ni-precursor. Subsequent removal
ous platinum catalysts, the reverse is observed, of the excess ammonia in flowing hydrogen at
cis-2-butene reacting faster than the trans-iso- 1808C gave a shape-selective catalyst over which
mer. 1-hexene was hydrogenated while 2,3-dimethyl-
In the same paper, a method for the selective 2-butene remained unconverted. Similar selec-
removal of ethene in propene streams is pre- tivity was obtained using a NirCa-LTA cata-
sented w14x. It is based on the sieving properties lyst, reduced in hydrogen at 4508C, from which
of Na-mordenite which can distinguish ethane the external surface located nickel was removed
from propane. The catalyst was made by ex- selectively by carbon monoxide at 808C. In this
changing platinumŽII. tetra-amine into the H- way, nickel on the external surface is removed
form of mordenite, and then narrowing the pore as NiŽ CO. 4 , a method which also proved to be
opening by sodium-ion back exchange. The cat- suitable for the Žpartial. regeneration of cata-
alyst was activated in dry air at 4508C and then lysts which had lost their shape selective proper-
treated with triphenylphosphine in hydrogen at ties as a result of severe temperature treatments.
2608C in order to poison selectively the plat- Dessau w17–19x developed the competitive
inum sites located on the external surface Ž ex- hydrogenation of an equimolar mixture of di-
ternal surface Pt.. Catalytic testing indicated the branched 4,4-dimethyl-1-hexene and linear 1-
selective conversion of ethene in a stream con- hexene as a method to assess the location of
taining predominantly propene. The observed platinum in PtrCs-ZSM-5 ŽSirAls 70.. Direct
selectivity might be explained by product selec- reduction of platinumŽII. tetra-amine exchanged
tivity w15x. Both propene and ethene can reach Cs-ZSM-5 in flowing hydrogen at 3008C gave a
the hydrogenationrdehydrogenation site; how- catalyst which did not show shape-selective
ever, ethane is the only hydrogenation product properties. Clearly, the platinum had migrated
that can leave the catalyst while propane re- to the external surface. Reduction of the catalyst
mains entrapped. Normally, this would result in in the presence of olefins, however, yielded a
deactivation of the catalyst as a result of the catalyst which showed a strong preference for
accumulation of propane. This was not ob- the reduction of the linear olefin. It was as-
served, however, which made the authors pro- sumed that the olefins reacted with the platinum
pose that equilibrium conditions might be estab- hydride species formed, which are responsible
lished in the interior of the zeolite between all for platinum migration and agglomeration. Oc-
50 E.J. Creyghton, R.S. Downingr Journal of Molecular Catalysis A: Chemical 134 (1998) 47–61

clusion of platinum clusters in the zeolite pores demonstrate the successful introduction of plat-
was also accomplished when the platinum-com- inum, palladium and rhodium in the small pore
plex loaded zeolite was first activated in an molecular sieves ZSM-58, Rho, ZK-5 and
oxygen stream at 3008C, prior to reduction in SAPO-42 by solid-state ion exchange with the
flowing hydrogen at the same temperature. The corresponding metal chlorides at elevated tem-
competitive hydrogenation of various other peratures w21x.
olefins was also investigated Ž Table 1. . Even A shape-selective catalyst for the selective
small PtrCs-ZSM-5 crystals, which have a rela- hydrogenation of acetylene in the presence of
tively large external surface, behaved in a butadiene and ethylene was developed by Corbin
shape-selective manner. Addition of a bulky et al. w22,23x. A rather fundamental study of the
phosphine, tri-p-tolylphosphine, further im- selective hydrogenation of acetylene impurities
proved the selectivity by poisoning the Pt lo- in ethene streams over palladium loaded zeolites
cated on the external surface. Repeated thermal had already shown the potential of catalysts of
treatment of the catalyst in hydrogen further this type in such reactions w24x. In the produc-
improved its shape-selective properties, which tion of ethene from steam cracking of natural
indicates that the occluded metal clusters are gas liquids Ž NGL. , small amounts of butadiene
stable, whereas external-surface platinum tends and acetylene are formed as by-products. Acety-
to agglomerate. lene can cause explosions in downstream cryo-
In addition to the test reaction discussed genic separation units and should therefore be
above, Weitkamp et al. w20x used a similar test removed from the gas stream. The recovery of
reaction in order to discriminate between inter- butadiene, however, is attractive from an eco-
crystalline platinum clusters in medium-pore ze- nomic point of view. On a conventional nickel
olites and platinum located on the external sur- catalyst, both acetylene and butadiene adsorb
face. The branched probe molecule 4,4-di- more strongly than ethylene so that minimal
methyl-1-hexene was replaced by the even more hydrogenation of ethylene is achieved while
branched molecule 2,4,4-trimethyl-1-pentene 60% of the butadiene and all the acetylene is
and used in competition with 1-hexene. It was converted.
successfully applied to illustrate the shape-selec- Zeolite LTA was chosen as support for a
tive hydrogenation properties of the medium more selective hydrogenation catalyst because
pore PtrZSM-5 and of the small pore, its pore opening of about 4.3 A˚ is in the range
PtrZSM-58. In the latter, platinum was intro- of the molecular dimensions of the reactants:
duced by co-crystallisation of platinumŽII. acetylene ; 3.3 A, ˚ ethylene ; 3.9 A˚ and buta-
tetra-amine which was added to the zeolite syn- diene ; 4.2 A. ˚ Furthermore, the pore opening
thesis mixture, as first reported by Weisz et al. of zeolite LTA can be fine-tuned by ion ex-
w12x. The test reaction was also applied to change with alkali or alkaline earth metal

Table 1
Shape-selective competitive hydrogenations over PtrCs-ZSM-5a . Data from Ref. w18x
Linear olefin Branched olefin Temperature Ž8C. % Hydrogenation
Linear Branched
Hexene-1 4,4-Dimethyl-1-hexene 275 90 1
Pentene-1 4,4-Dimethyl-1-pentene 300 97 2
Heptene-1 4,4-Dimethyl-1-pentene 300 91 1
Hexene-1 6-Methyl-1-heptene 300 25 2
Styrene 2-Methylstyrene 425 50 2
a
Conditions: down-flow fixed-bed glass reactor; equimolar mixture of olefins.
E.J. Creyghton, R.S. Downingr Journal of Molecular Catalysis A: Chemical 134 (1998) 47–61 51

cations. Nickel ions were introduced by ion non-acidic tin-, lead-, or indium-modified
exchange with aqueous nickel nitrate. Reduction PtrZSM-5; typical reactants which can enter
of the metal ions was performed in hydrogen at this zeolites comprise benzene, toluene and
4008C. The accessibility of the ion-exchange xylenes, as well as other monoalkylbenzenes
zeolites was investigated by sorption studies and b-alkylnaphthalenes. Bulkier aromatics are
performed with the probe molecules MeOH Ž; excluded and thus remain unconverted. The
3.7 A ˚ ., EtOH Ž; 4.0 A˚ . and n-PrOH Ž; 4.3 modified PtrZSM-5 catalysts showed shape-
A˚ ., whose molecular dimensions are just slightly selective hydrogenation of para-xylene, yield-
larger than those of the actual reactant ing 1,4-dimethylcyclohexane, when applied in
molecules. No n-PrOH was adsorbed by Kq-, the competitive hydrogenation of an equimolar
Rbq- and Csq-exchanged zeolite LTA, which mixture of ortho- and para-xylene at 3258C;
indicates that these catalysts should show re- selectivities up to 47 pararortho were reported.
duced hydrogenation of butadiene. In addition, Furthermore, in the competitive hydrogenation
for Cs-LTA strongly reduced adsorption of of an equimolar mixture of benzene, toluene
EtOH was found, which means that ethylene, and para-xylene at 2508C the order of reactiv-
formed by hydrogenation of acetylene, might be ity was found to be benzene) toluene) para-
unable to leave the Csq-exchanged catalyst. xylene which is in contrast with the order found
Initially, no shape selectivity was observed for a sulfided NirW on alumina catalyst.
for any of the zeolite-based catalysts, indicating
that the reaction occurred over nickel sites lo- 2.2. Liquid phase reactions
cated on the external surface. The presence of
external-surface nickel was confirmed by ESCA Huang and Schwartz w26x have prepared
measurements. Poisoning of these sites by phos- rhodium complexes within the zeolite cavities
phines or thiophene, or modification with te- of partially proton-exchanged Linde 13 = type
traethyl orthosilicate ŽTEOS. or leadŽ II. 2-ethyl- molecular sieve ŽNa-X, FAU structure.. Reac-
hexanoate proved successful in conferring shape tion of triallylrhodium with Brønsted acid sites,
selectivity. Only 20% butadiene loss was found located in the supercages, gave the supported
at full acetylene conversion over NirK-LTA diallylrhodium complex 1 ŽScheme 1.. Subse-
ŽthiophenerH 2 S added. while using conditions quent treatment of 1 with hydrogen led to the
close to those commonly applied in industry. A formation of the zeolite-bound rhodium hydride
gradual decrease in activity made regeneration 2 and the evolution of one equivalent of propane.
necessary after 30 days. Calcination in air at No bridging hydride ligands were detected by
4508C restored the activity; however, selectivity IR analysis, which indicated that the complexes
was decreased by about 5% per cycle. This was were monomerically distributed over the zeolite.
attributed to a resiting of the nickel.
Recently, Dessau w25x reported a shape-selec-
tive hydrogenation process which is capable of
selectively removing certain aromatic com-
pounds from a mixture with others. Such a
process is needed, from an environmental point
of view, for the reduction of the aromatic con-
tent Ž especially benzene. in gasoline. Further-
more, a decrease in the amount of aromatics in
a refinery distillate employed in the production
of kerosine results in a decrease in soot produc- Scheme 1. Formation of zeolite-encapsulated rhodium-hydride 2,
tion during combustion. The catalyst consists of according to Ref. w26x.
52 E.J. Creyghton, R.S. Downingr Journal of Molecular Catalysis A: Chemical 134 (1998) 47–61

The zeolite-encapsulated hydride 2 showed a mixture did not change the hydrogenation rate
linear uptake of hydrogen in the catalytic hydro- of 1-butene over 2. Addition of excess PMe 3 ,
genation of all olefins studied, comparable to however, caused the hydrogenation rate to de-
the behaviour of its silica-supported analogue. crease to ca. zero.
Negligible hydrogenation rates, however, were Yamaguchi et al. w27x and Joh et al. w28x
found for substrates larger than cyclohexene reported the shape-selective hydrogenation of
Že.g., cyclooctene. , which clearly demonstrates olefins by a rhodium zeolite-Y catalyst. The
the sieving properties of the catalyst. Further- starting material was a rhodiumŽIII. exchanged
more, it proves that the active hydrogenation Na-Y zeolite ŽFAU structure., wRhŽNH 3 . 6 x 3qr
centres are located within the micropores of the Na-Y. Calcination of the dry material Ž 4008C. ,
zeolite-based catalyst. Whereas the hydrogena- followed by treatment with wet carbon monox-
tion rates for 1-butene, 1-hexene and 1-octene ide Ž1308C. gave pink rhodium carbonyl clus-
over silica-supported rhodium hydride com- ters Rh mŽ CO . nrNa-Y. The formation of
plexes are rather similar, and greater than that rhodium carbonyl clusters was confirmed by IR
of the branched olefin 2,3-dimethyl-2-butene, spectroscopy. Subsequent decarbonylation with
this is not true for the zeolite-bound complex. hydrogen Ž1208C. gave the active Rh xrNa-Y
Over catalyst 2 the rates were found to decrease catalyst. Shape-selective hydrogenations of vari-
in the order 1-butene ) 1-hexene ) 2,3- ous olefins were performed without poisoning
dimethyl-2-butene) 1-octene, which was as- of rhodium clusters possibly present on the
cribed by the authors to transport limitations external surface. Relative hydrogenation rates
within the zeolite cavities. Selective poisoning were found to decrease in the order 1-hexenef
experiments further confirmed that the active 1-octene f cyclohexene ) cyclooctene 4 cyc-
hydrogenation sites were located within the zeo- lododecene. Competitive hydrogenation experi-
lite matrix. Coordination of two equivalents of ments using a mixture of cyclohexene and cy-
PMe 3 to the zeolite-bound rhodium complex 2 clododecene showed that only cyclohexene was
yields 3, while n-Bu 3 P because of its larger hydrogenated over the zeolite-based catalyst
cone angle is not able to coordinate with the while both substrates were converted over a
noble metal site within the zeolite structure. RhrC reference catalyst ŽFig. 1. . These results
Addition of excess of n-Bu 3 P to the reaction indicate the presence of the hydrogenation sites

Fig. 1. Competitive hydrogenation of cyclohexene Ž`. and cyclododecene Žv . Žconversion.: substrate 5 mmol each; hexane 20 ml; 101 kPa
hydrogen; Ža. Rh xrNa-Y, 50 mg, 508C; Žb. 5% RhrC, 20 mg, 308C. Redrawn using data from Ref. w27x.
E.J. Creyghton, R.S. Downingr Journal of Molecular Catalysis A: Chemical 134 (1998) 47–61 53

within the zeolite cavities. Furthermore, they non-selective. Furthermore, because of their ac-
indicate the potential of noble metal in zeolite cessibility, they might dominate the overall re-
systems as shape-selective hydrogenation cata- action selectivity. Selective poisoning of these
lysts. surface-located rhodium clusters with n-Bu 3 P,
Corbin et al. w29x prepared various rhodium- as reported by Huang and Schwartz w26x, was
containing zeolites and tested them as catalysts found to increase the cyclopentener4-methyl-
in the competitive hydrogenation of cyclopen- cyclohexene selectivity by a factor 4 in case of
tene and 4-methylcyclohexene in the liquid RhrZSM-11. All catalysts were therefore
phase. Rhodium was either introduced by the treated with n-Bu 3 P in order to maximize selec-
method of Huang and Schwartz Žintroduction of tivity. The second factor involved the variable
trisŽp-allyl.rhodium w26x. or by ion-exchanging water content. For RhrX the selectivity was
the zeolites with aqueous rhodiumŽ III. chloride. found to increase from 1 to 30 with increasing
The zeolites used included the small-pore ZSM- water content Ž2 to 32 wt.%., while RhrZSM-11
34, the medium-pore ZSM-5 and ZSM-11, and exhibited a maximum of 47 at intermediate
the large-pore zeolites mordenite, X and Y. water content. This effect was explained as a
Sorption measurements were performed on the tuning of the intracrystalline pore dimensions by
dehydrated zeolites with the probe molecules means of a water coating, lining the zeolite
cyclopentane and cyclohexane, having dimen- pores and cages. Moreover, the hydrophilic
sions comparable to those of the reactants character of the zeolite catalysts will increase. It
Žthough methylcyclohexane would have been a should be noted, however, that the RhrX zeo-
better analogue for 4-methylcyclohexene. . These lite with 32 wt.% water is completely saturated
indicated hardly any intrinsic size selectivity of and can hardly accommodate additional organ-
the zeolite framework itself and in accordance ics. Probably the activity originates from some
with this, no correlation was observed between non-deactivated Rh on the external surface of
the pore size of the zeolites and the reaction the zeolite crystals.
selectivity in the hydrogenation experiments A solid catalyst for the shape-selective hy-
Table 2. Moreover, two additional factors, apart drogenation of molecules with a diameter greater
from pore-size differences, appeared to have a than about 9 A ˚ was reported by Davis w30x. It
major impact on the reaction selectivity. The consists of an aluminophosphate molecular
first was related to the location of the noble sieve, VPI-5, which has been impregnated with
metal sites. Hydrogenation sites located on the a noble metal salt solution and subsequently
external surface were expected to be essentially reduced. The catalyst is claimed to show

Table 2
Selectivity in the competitive hydrogenation of cyclopentener4-methylcyclohexene over Rh-containing zeolitesa . Data from Ref. w29x
Zeolite ˚.
Pore size ŽA Sorption selectivity b Reaction selectivity c Water content Žwt.%. d
ZSM-34 3.6 = 5.1 0.77 7.5 y
ZSM-5 5.3 = 5.6 1.54 15.0 1.2
ZSM-11 5.3 = 5.4 1.10 47.0 2.2
Mordenite 6.5 = 7.0 0.91 16.0 8.2
X 7.4 0.91 6.0 5.0
Y 7.4 0.96 1.0 3.6
a
Conditions: 30 mg Rh-zeolite, 65 ml n-octane, 0.3 ml cyclopentene, 0.3 ml 4-methylcyclohexene and 5 m l PBu 3 at 608C and 345 kPa
hydrogen.
b
Sorption selectivity of the anhydrous zeolite for cyclopentanercyclohexane.
c
Selectivity over the partially hydrated Žsee next column. Rh-zeolites in the competitive hydrogenation of cyclopentener4-methylcyclohe-
xene.
d
Water content that gave rise to the presented reaction selectivity.
54 E.J. Creyghton, R.S. Downingr Journal of Molecular Catalysis A: Chemical 134 (1998) 47–61

shape-selective properties in the hydrogenation catalytic activity and regioselectivity were only
of olefins, aldehydes and ketones, provided that obtained when non-polar or bulky solvents were
external surface Pt is poisoned by bulky phos- applied. The hydrogenation rate over PtrNa-
phines or mercaptans. The activity of a BEA was found to be much lower than that
RhrVPI-5 catalyst in the hydrogenation of observed over the amorphous supported plat-
molecules larger than about 9 A ˚ was illustrated inum catalysts. Furthermore, for the zeolite-
by the liquid-phase hydrogenation of a 1:1 based catalyst, first-order kinetics were found
Žwrw. mixture of 1-octene and cis-cyclooctene, and compared with zero-order for the amor-
performed in decane at 308C and 101 kPa hy- phous catalysts. Both effects were explained in
drogen. Both reactants were converted over terms of alkene coverage of the Pt-sites in the
RhrVPI-5 while only the linear olefin was con- two types of catalysts.
verted over the large-pore molecular sieve
RhrAlPO4-5.
Recently, Creyghton et al. w31x reported on 3. Chemoselective hydrogenation of a , b-un-
the liquid-phase regioselective hydrogenation of saturated aldehydes over zeolite-encapsulated
alkenes over Pt-loaded zeolite BEA. It was found metal clusters
that 1-decene was hydrogenated 18 times faster
than trans-5-decene over PtrNa-BEA, whereas The chemoselective hydrogenation of cin-
the ratio was only about 2 for platinum on namaldehyde 4 to cinnamyl alcohol 5 over plat-
non-microporous supports. This regioselectivity inum and rhodium clusters encapsulated in Y-
was explained by steric constraints imposed by type zeolite was studied by Gallezot et al.
the microporous structure of the zeolite. Careful ŽScheme 2. w32x. From a thermodynamic point
study of the possible adsorption sites on a zeo- of view, both the hydrogenation of the C5C
lite-occluded platinum cluster indicated more and the C5O bond are possible but the reaction
possibilities for the adsorption of a terminal enthalpy ŽyD H . for the hydrogenation of the
double bond, as compared to an internal double C5C bond is about 42 kJrmol higher. More-
bond. However, regioselectivity was only ob- over, the selective hydrogenation of the C5C
tained when the non-selective, external-surface bond, to give 3-phenylpropanal 6, is rather triv-
located platinum clusters were selectively poi- ial, while it is much more difficult to selectively
soned with triphenylphosphine. Furthermore, convert the carbonyl in order to obtain the
sorption experiments showed that strong compe- unsaturated alcohol 5. Three different catalysts
tition between the solvent and the reactant for were prepared by Gallezot et al. w32x; PtrYŽ1.,
adsorption in the zeolite might exist, which PtrYŽ2. and RhrY. The platinum clusters in
prevents the reactants from attaining a signifi- PtrYŽ1. were homogeneously distributed over
cant concentration in the pores. As a result, the zeolite matrix and their sizes were in the

Scheme 2. Hydrogenation of cinnamaldehyde 4 to: cinnamyl alcohol 5, 3-phenylpropanal 6 and 3-phenyl-1-propanol 7.


E.J. Creyghton, R.S. Downingr Journal of Molecular Catalysis A: Chemical 134 (1998) 47–61 55

range 1–2 nm. In contrast, PtrYŽ2. contained higher basicity of the K-exchanged zeolite. The
platinum clusters of about 50 nm, located in the suggested enhancement of electron density thus
vicinity of the external surface but still within induced on the metal clusters was assumed to
the matrix. RhrY contained small clusters in result in a decreased C5C hydrogenation activ-
the range of 1 " 0.3 nm, homogeneously dis- ity, as well as in an increase in C5O hydro-
tributed throughout the zeolite crystal. The hy- genation caused by the interaction of the car-
drogenation reactions were performed in a bonyl with the zeolitic cation.
stirred autoclave at 608C and 40 bar hydrogen. Gallezot et al. w34x continued their investiga-
At a conversion of 50%, the selectivities to 5 tion of the regioselective hydrogenation of 4 to
were found to be 82 and 97% for PtrYŽ1. and 5 over zeolite PtrBEA. Two catalysts were
PtrYŽ2., respectively, while the selectivity for a prepared, differing in metal cluster size and
PtrC reference catalyst was only 55%. Simi- location. PtrNa-BEAŽ1. was prepared by direct
larly, 30% selectivity was found for RhrY, reduction of the platinumŽ II. tetra-amine ex-
which is relatively high for a rhodium catalyst changed zeolite, by heating in flowing hydrogen
as this metal is known to be very unselective. from 25 to 2008C at 28Crmin, while PtrNa-
Moreover, only 5% selectivity was observed for BEAŽ2. was obtained by calcination from 25 to
a RhrC reference catalyst. These unusually high 3008C at 18Crmin in flowing oxygen, cooling
selectivities were explained by geometric con- under argon to 1008C and heating under flowing
straints imposed by the zeolite structure. Be- hydrogen from 100 to 3008C. The PtrNa-
cause of its rigidity, 4 can only adsorb end-on BEAŽ1. contained zeolite-encapsulated grape-
across the 12-membered oxygen ring that gives like agglomerates of 50–200 nm, which con-
access to the supercage-encapsulated metal clus- sisted of metal crystallites of 2–5 nm, compara-
ters. This argument also holds for clusters larger ble to PtrYŽ2. w14x. The PtrNa-BEAŽ2. con-
than the supercage which is still claimed to be tained small platinum clusters with size compa-
located within the matrix. Small metal clusters, rable to the pore opening Ž- 1 nm., homoge-
however, which contain only a few atoms, would neously distributed over the zeolite matrix. A
leave enough space in the supercages for 4 to high selectivity to 5 was obtained Ž86–88%.
adsorb laterally via the C5C bond. This could over PtrNa-BEAŽ1., independent of the conver-
well be the case for PtrYŽ 1. and RhrY. For the sion. As in the case of PtrYŽ 2. w32x, this was
latter, external-surface located rhodium parti- explained by a hindered approach of the C5C
cles, which have a high activity for C5C hydro- bond; end-on adsorption of the C5O is favoured
genation, were also considered. when the molecule diffuses parallel to the linear
This research was extended to RurNa-Y and channel wall, until it reaches the encapsulated
RurK-Y which also gave improved selectivities platinum cluster ŽFig. 2..
to 5 as compared with RurC w33x. In order to In contrast, over PtrNa-BEAŽ2. a low initial
investigate the importance of steric considera- selectivity Ž0–5%. for the unsaturated alcohol
tions, an aliphatic molecule, 3-methyl- was obtained, while 3-phenyl-1-propanol 7 was
crotonaldehyde, was also included. This less the main product. After 1 h, however, the for-
bulky molecule has more space to adsorb in mation of 5 became predominant Ž50–55%.
many orientations in the confined interior of the which was accompanied by a decrease in hydro-
zeolite; steric effects were therefore suggested genation rate. It was proposed that the reaction
to be less important. The selectivity to the un- initially occurred over metal clusters of low
saturated alcohol could be increased, however, nuclearity, which leave enough space between
by replacement of the Na-ions by K-ions. This the cluster and the zeolite wall to allow the
was explained by Blackmond et al. w33x to be lateral adsorption of 4, resulting in the hydro-
the result of an electronic effect, related to the genation of both double bonds. During reaction,
56 E.J. Creyghton, R.S. Downingr Journal of Molecular Catalysis A: Chemical 134 (1998) 47–61

related to the zeolite structure, play a minor role


as evidenced by the predominant formation of
the saturated aldehyde over PdrK-LTL. More-
over, liquid-phase reactions in the one-dimen-
sional channel system of zeolite LTL are ex-
pected to be slow as a result of diffusional
limitations.

4. Shape-selective hydrogenations of olefins


over zeolite-occluded coordination com-
pounds; ship-in-a-bottle systems
Fig. 2. Cinnamaldehyde approaching a metal cluster; Ža. pore
filling cluster; Žb. low-nuclearity cluster, allowing lateral adsorp- The encapsulation of metalrligand com-
tion. Illustration adapted from Ref. w34x. plexes in zeolites combines the catalytic proper-
ties of the coordination complex with the
shape-selective properties of the molecular sieve.
the number of these clusters decreases by sinter- Moreover, the heterogenised catalysts offer the
ing and the reaction continues over a smaller benefit of easy catalyst separation. Furthermore,
number of pore-filling platinum clusters where possible deactivation by complex oligomerisa-
steric constraints predominate. Recycle experi- tion is prevented. So far, most papers on ship-
ments showed that the initial selectivity of the in-a-bottle systems focus on selective oxidation
second run was identical with the selectivity catalysis, while the number of papers dealing
during the second part of the first run, which with the application of ship-in-a-bottle com-
supports this hypothesis. plexes in selective catalytic hydrogenation is
Li et al. w35x studied the chemoselective hy- still very limited.
drogenation of 4 over PtrL ŽM s alkali or alka- Kowalak et al. w36x reported on the catalytic
line earth metal cation.. High selectivities to 5 properties of PdŽ II.Ž salen. Ž Fig. 3. encapsulated
Ž) 90%. were obtained over rubidium- and in the supercages of zeolites Y and X. Upon
strontium-exchanged catalysts at conversions up hydrogenation of 1-hexene at room temperature
to 95%. By X-ray powder diffraction some and 54.3 kPa hydrogen Žliquid phase., signifi-
characteristic peaks for platinum were identified cant isomerisation to trans-2-hexene was ob-
on the PtrM-LTL catalysts, which indicates served, which complicated the evaluation of
that the clusters are rather large Ž) 2.5 nm. . shape selectivity based on single-component hy-
Presumably, these large clusters are located on drogenation rate measurements for the reactants
the external surface, in contrast to the situation 1-hexene and trans-2-hexene. Furthermore, cy-
with the catalysts described by Gallezot et al. clohexene was converted solely to cyclohexane
w32,34x and Blackmond et al. w33x, while some over both homogeneous and encapsulated
small clusters might be located inside the micro- PdŽsalen. while for palladium-exchanged zeo-
pores; further characterisation Že.g., TEM. might lites dehydrogenation to benzene was also ob-
confirm this hypothesis. The high selectivity to
the unsaturated alcohol is controlled by several
factors including the nature of the noble metal
ŽPt ‘favours’ C5O hydrogenation., the metal
morphology, and electronic effects induced by
the basicity of the zeolite. Steric considerations, Fig. 3. PdŽII.Žsalen..
E.J. Creyghton, R.S. Downingr Journal of Molecular Catalysis A: Chemical 134 (1998) 47–61 57

served. The latter result was attributed to the


formation of PdŽ0. in the zeolite. In a competi-
tive hydrogenation experiment, applying a 1:1
Žvrv. mixture of 1-hexene and cyclohexene, the
zeolite-encapsulated PdŽ salen. produced only
hexane and trans-2-hexene while cyclohexene Fig. 4. The immobilised chiral rhodium complex applied by
was not converted. No preference was observed Corma et al. for the asymmetric hydrogenation of prochiral
alkenes. Illustration adapted from Ref. w38x.
for the homogeneous PdŽsalen. complex. The
preferential conversion of the terminal double
bond was also observed in a 20r1 Žvrv. cyclo-
hexener1-hexene mixture. Apparently, the was made by steam calcination of 80% ammo-
presence of a terminal double bond which is nium-exchanged Na-Y and contained a well
sterically less demanding largely denies the developed micro-rmesoporous channel system
cyclic alkene access to the supercage. Alterna- with pore diameters between 12 and 30 A. ˚ Both
tively, the coordination of cyclohexene to su- the homogeneous and the immobilised chiral
percage located PdŽII.Žsalen. might be strongly complexes were tested as catalysts in the
hindered because of steric considerations. liquid-phase hydrogenation of prochiral alkenes
The hydrogenation of cyclohexene, cy- Ž a-acylaminocinnamate derivatives. . Whereas
clooctene, 1-hexene and benzene by NiŽsalen. induction periods were observed for both the
encapsulated in zeolite Y, performed in the homogeneous and silica-anchored catalysts, this
liquid phase at 408C and 60 atm Ž 1 atm s was not the case for the USY-supported com-
101.325 kPa. hydrogen, was studied by Chatter- plex. This result was explained by a concentra-
jee et al. w37x. All the olefins and the benzene tion effect, occurring in the pores of the zeolite,
were successfully converted to their hydro- andror the specific electrostatic interaction be-
genated products. The mechanism was proposed tween the substrate and the zeolitic support.
to involve the formation of a high-valent hy- Furthermore, the USY-anchored catalyst gave
drido nickel complex which transfers hydrogen the highest activity in combination with in-
atoms to the substrates. In the competitive hy- creased enantiomeric excess Žee.. The improved
drogenation of a 1r1 Žvrv. mixture of 1- ee was most pronounced for the less bulky
hexenercyclohexene, n-hexane was observed substrates, suggesting an Žimportant. role for
as the only product. The terminal double bond, steric constraints imposed by the zeolite struc-
which is sterically more accessible, was be- ture. Later, similar results were reported for
lieved to prevent the cyclic alkene from coordi- comparable USY-supported rhodium and nickel
nating to the catalytic centre. complexes w39x.
The enantioselective hydrogenation of ethyl
pyruvate over Žy.-cinchonidine-modified plat-
5. Noble metal containing zeolites as enan- inum, introduced in zeolite Y, ZSM-5, ZSM-35
tioselective hydrogenation catalysts and BEA, was studied by Reschetilowski et al.
w40x. The highest activity and enantioselectivity
Enantioselective hydrogenation reactions are was obtained for the ZSM-35 based catalyst.
of increasing importance in the pharmaceuticals However, enantiomeric excess was also ob-
and agrochemicals industry. Recently, Corma et served to be strongly influenced by the type of
al. w38x synthesised Rh-complexes with chiral solvent. Unfortunately, the specific benefit
N-containing ligands, derived from L-proline. andror influence of the zeolite support on the
These complexes were successfully anchored on enantioselectivity was not discussed. Additional
silica and zeolite USY ŽFig. 4.. The USY-zeolite work can be found in Refs. w41,42x.
58 E.J. Creyghton, R.S. Downingr Journal of Molecular Catalysis A: Chemical 134 (1998) 47–61

6. Shape-selective hydrogenation of olefins value of 9. Although these experiments could


over zeolite-containing composite catalysts not exclude the presence of mesopores in the
zeolite layer, they did show the specific influ-
A novel type of hydrogenation catalyst con- ence of the zeolite micropores. An interesting
sisting of a TiO 2rPtrsilicalite-1 composite was observation was the more pronounced double-
recently reported by van der Puil et al. w43,44x. bond migration in the case of the composite
The TiO 2 support offers high stability for the catalyst, as compared to uncovered PtrTiO 2 .
catalytic phase Žplatinum. while a thin continu- Because of the absence of acid sites on the
ous layer of oriented silicalite crystals, which monofunctional catalyst system, it was proposed
covers the active sites, creates separation selec- that the hydrogen-to-olefin ratio at the catalytic
tivity and steric constraints at the zeoliter site of the composite was relatively low as a
platinum interface ŽFig. 5. . The composite was result of a retarded hydrogen supply through the
made by sputtering of platinum onto one side of zeolite layer. In the competitive hydrogenation
a TiO 2 single crystal platelet, followed by the of 1-decene and trans-5-decene, preferential hy-
hydrothermal synthesis of a silicalite-1 layer. drogenation of the terminal olefin was observed.
This resulted in a monofunctional catalyst which This regioselectivity was explained by steric
contained platinum clusters with a diameter of constraints present at the catalytic site w31x.
5–20 nm, distributed homogeneously over the Moreover, this indicates a close contact between
support, and covered with a silicalite-1 layer the zeolite and the catalytic phase at the zeo-
with a thickness of about 1 m m. Careful re- literplatinumrTiO2 interface.
moval of the organic template by calcination
Ž3508C. and subsequent treatment in ozone
Ž1208C. cleared the access to the platinum sites 7. Zeolite-catalysed Meerwein–Ponndorf–
through the micropores while minimizing the Verley (MPV) reductions
formation of mesoporesrcracks. In the competi-
tive gas-phase hydrogenation of 1-heptene and Shabtai et al. w45x studied the potential of
3,3-dimethyl-1-butene, performed at 1008C, a alkali or alkaline earth exchanged X-type zeo-
maximum hydrogenation selectivity of 70 in lites in the gas-phase Ž 100–1808C. MPV reduc-
favour of the conversion of the linear olefin was tion of various saturated and unsaturated alde-
found. This was attributed to diffusional limita- hydes and ketones, using isopropanol as reduc-
tion, resulting in mass-transport selectivity. ing agent w45x. In the reduction of linear aldehy-
Upon raising the reaction temperature to 2008C, des over Na-X, a gradual decrease in the reduc-
the influence of diffusion limitations decreased. tion rate was observed upon increasing chain
As a result, the conversion rates of both olefins length, which was attributed to increasing diffu-
increased while the selectivity decreased to a sional limitations in the micropores. Selectivi-
ties to the corresponding 1-alcohols were gener-
ally high Ž) 95%.. Application of Lewis-acidic
Ca-X gave acetalisation of the aldehydes as an
important side-reaction. This could be pre-
vented, however, by applying higher reaction
temperatures. Unfortunately, the X-type zeo-
literisopropanol system was not capable of re-
ducing a , b-unsaturated aldehydes. Shape selec-
tivity was found in the selective conversion of
Fig. 5. Schematic representation of the TiO 2 rPtrsilicalite-1 com- citronellal 8 under MPV conditions. In Na-X
posite catalyst. Illustration adapted from Ref. w44x. there was enough space for the substrate to
E.J. Creyghton, R.S. Downingr Journal of Molecular Catalysis A: Chemical 134 (1998) 47–61 59

27
perform an intramolecular ring closure to iso- Al-NMR spectra did not show any increase in
pulegol 9 whereas, over Cs-X, reduction to the octahedral aluminium. FT-IR results indicated a
linear citronellol 10 was observed ŽScheme 3.. relation between the catalytic activity and the
In the reduction of methylcyclohexanone iso- amount of aluminium which is only partially
mers Ž1008C., it was observed that the 4-isomer bonded to the framework ŽLewis-acid sites. .
reacted relatively fast and gave a thermodynam- The MPV mechanism was therefore proposed to
ically determined product distribution Žcis:trans involve a six-membered transition state, analo-
s 24:76.. The 2- and 3-methylcyclohexanone gous to that of AlŽOPr i . 3-catalysts, formed upon
reacted more slowly and gave a kinetically de- chemisorption of a secondary alcohol on a
termined product distribution Ž cis:trans s Lewis-acid aluminium site and coordination of
62.5:37.5 and 23.5:76.5 for the 2- and 3-isomer, the ketone to the same site. A base mechanism
respectively.. The mechanism was proposed to was ruled out because of the low aluminium
involve the formation of a surface isopropoxide content Ž SirAls 12., the absence of alkali or
group attached to a cationic site Žbasic mecha- alkaline earth cations in the active H-BEA cata-
nism.. It could not be excluded, however, that lysts and the very similar activity of the Li-,
incompletely coordinated Si- or Al-sites con- Na-, K-, Rb- and Cs-exchanged catalysts. Fur-
tributed to the catalytic activity Ž Lewis-acid thermore, the catalyst could be poisoned by the
mechanism. . base piperidine.
Recently, Creyghton et al. w46,47x reported The transition states which lead to the cis- or
the application of zeolite beta ŽBEA. in the trans-alcohol differ substantially in spatial re-
stereoselective Ž) 95%. reduction of 4-tert- quirements Ž Fig. 6.. That for the cis-isomer is
butylcyclohexanone to cis-4-tert-butyl-cyclo- more or less linear in form and aligned with the
hexanol in the liquid phase. This zeolite-based BEA channel, while the formation of the trans-
catalyst proved to be fully regenerable without alcohol requires an axially oriented Ž bulkier.
loss in activity or stereoselectivity. This is of transition state Žfor spatial pictures, the reader is
industrial relevance, as the cis-isomer is a fra- referred to Fig. 8 of Ref. w47x.. Although the
grance-chemical intermediate. Other active solid latter might still fit in the intersections of BEA
catalysts, including zeolites, invariably gave the it is questionable whether there is an active site
thermodynamically more stable trans-isomer. available at the required position. More coordi-
The activity of the BEA catalyst was found to nation possibilities are available for the cis-tran-
increase upon increasing activation temperature. sition state, which can easily be accommodated
Furthermore, deep-bed calcination conditions within the straight channels of BEA. The ob-
gave a higher catalytic activity than a shallow- served kinetically determined product distribu-
bed procedure, indicating a relation between the tion is thus satisfactorily explained by true tran-
catalytic activity and the extent of framework sition-state selectivity.
dealumination since the former method results In an extension of this work, van der Waal et
in a greater degree of auto-steaming. However, al. w48x reported the catalytic activity of alu-
minium-free titanium beta ŽwTix-BEA. zeolite in
the same stereoselective MPV reaction. Again, a
very high selectivity of 98% to the cis-isomer
was found, explained by a restricted transition
state, in this case around a Lewis-acid titanium
site. The Lewis-acid properties of tetrahedrally
incorporated titanium in zeolite wTix-BEA had
Scheme 3. Shape selective conversions of citronellal 8 to isopule-
gol 9 or citronellol 10 under MPV conditions, according to Ref. already become clear during catalytic studies on
w45x. the epoxidation of olefins with hydrogen perox-
60 E.J. Creyghton, R.S. Downingr Journal of Molecular Catalysis A: Chemical 134 (1998) 47–61

metal clusters or coordination complexes have


been reported. Most of them exemplify the siev-
ing properties of these catalysts, resulting in the
selective hydrogenation of those substrates
which can enter the micropores, while leaving
the larger substrates unaffected. In order to ob-
serve and maximise this reactant shape selectiv-
ity, metal clusters located on the external sur-
face should be deactivated by selective poison-
ing.
A new approach involves the construction of
zeolite-containing composite catalysts, in which
supported metal clusters are covered with a thin,
oriented zeolite layer. The support offers high
stability for the catalytic phase, while the zeolite
creates selectivity by its separation capacity and
steric constraints created at the catalytic sites, as
first illustrated for a TiO 2rplatinumrsilicalite-1
Fig. 6. Transition states for the formation of cis-4-tert-butyl-
composite.
cyclohexanol Žtop. and trans-4-tert-butylcyclohexanol Žbottom. in For chemo-, regio- or stereoselectivity in a
the proposed reaction mechanism over zeolite BEA. Illustration single molecule, only a few examples appear to
from Ref. w47x.
be known. They clearly demonstrate the poten-
tial of metal-loaded zeolites as highly selective
ide in alcoholic solvents w49x. The oxophilic hydrogenation catalysts. However, the factors
Lewis acidity of the titanium site was confirmed that govern selectivity and activity are not yet
by UV–Vis spectroscopy, which showed an fully elucidated and may include, in addition to
increased coordination number for the originally competitive adsorption of reactants, products
4-coordinated titanium atom upon adsorption of and solvents, also steric and electronic effects.
alcohols and water. Kinetically determined Moreover, further improvement in our under-
product distributions were also obtained in the standing of these catalytic phenomena is re-
MPV reduction of 2-, 3- and 4-methylcyclo- quired in order to make optimal use of the
hexanone; the cis-, trans- and cis-alcohols, re- possibilities which metal-in-zeolite systems cer-
spectively, being the major products. The cat- tainly offer.
alytic activity of wTix-BEA was found to be The recognition of the utility of zeolites as
much lower than that of its aluminium ana- catalysts in the MPV reduction of carbonyl
logue, although its tolerance for water was ob- compounds is very recent. Their potential is
served to be much higher. The latter property, related to the possibility of shape-selectivity, as
which is related to the hydrophobic character of illustrated by an example showing almost abso-
the aluminium-free zeolite, illustrates its cat- lute stereoselectivity as a result of restricted
alytic potential for this type of reaction. transition-state selectivity. In the case of alkali
or alkaline earth exchanged zeolites with a high
aluminium content ŽX-type., the catalytic activ-
8. Concluding remarks ity is most likely related to basic properties. For
zeolite BEA ŽSirAl) 10., however, aluminium
Many examples of shape-selective hydro- atoms which are only partially connected to the
genation reactions using zeolite-encapsulated framework appear to play a role in the catalytic
E.J. Creyghton, R.S. Downingr Journal of Molecular Catalysis A: Chemical 134 (1998) 47–61 61

activity. Similarly, the activity of aluminium- w22x D.R. Corbin, C. Bonifaz, L. Abrams, Stud. Surf. Sci. Catal.
38 Ž1988. 295.
free wTix-BEA in the MPV-type hydrogen-trans- w23x D.R. Corbin, L. Abrams, C. Bonifaz, J. Catal. 115 Ž1989.
fer reaction is ascribed to the Lewis-acid charac- 420.
ter of the titanium atoms. w24x R.P. Denkewicz, A.H. Weiss, W.L. Kranich, J. Wash. Acad.
Sci. 74 Ž1. Ž1984. 19.
w25x R.M. Dessau, US Pat., 5.350.504, 1994.
w26x T.-N. Huang, J. Schwartz, J. Am. Chem. Soc. 104 Ž1982.
Acknowledgements 5244.
w27x I. Yamaguchi, T. Joh, S. Takahashi, J. Chem. Soc., Chem.
The authors would like to express their ap- Commun. Ž1986. 1412.
w28x T. Joh, I. Yamaguchi, S. Takahashi, Nippon Kagaku Kaishi,
preciation to Professor Herman van Bekkum for 1989, p. 487.
his inspiration and unfailing enthusiasm. w29x D.R. Corbin, W.C. Seidel, L. Abrams, N. Herron, G.D.
Stucky, C.A. Tolman, Inorg. Chem. 24 Ž1985. 1800.
w30x M.E. Davis, US Pat., 5.406.012, 1995.
w31x E.J. Creyghton, R.A.W. Grotenbreg, R.S. Downing, H. van
References Bekkum, J. Chem. Soc., Faraday Trans. 92 Ž1996. 871.
w32x P. Gallezot, A. Giroir-Fendler, D. Richard, Catal. Lett. 5
w1x R.L. Augustine, Heterogeneous Catalysis for the Synthetic Ž1990. 169.
Chemist, Marcel Dekker, New York, 1996, p. 447. w33x D.G. Blackmond, R. Oukaci, B. Blanc, P. Gallezot, J. Catal.
w2x M.M.J. Treacy, J.M. Newsam, Nature 332 Ž1988. 249. 131 Ž1991. 401.
w3x J.M. Newsam, M.M.J. Treacy, W.T. Koetsier, C.B. De- w34x P. Gallezot, B. Blanc, D. Barthomeuf, M.I. Paıs ¨ da Silva,
Gruyter, Proc. R. Soc. London A 420 Ž1988. 375. Stud. Surf. Sci. Catal. B 84 Ž1994. 1433.
w4x M.M.J. Treacy, J.M. Newsam, Nature 332 Ž1988. 249. w35x G. Li, T. Li, Y. Xu, J. Chem. Soc., Chem. Commun. Ž1996.
w5x A.P. Stevens, P.A. Cox, J. Chem. Soc., Chem. Commun. 497.
Ž1996. 343. w36x S. Kowalak, R.C. Weiss, K.J. Balkus, Jr., J. Chem. Soc.,
w6x M.E. Davis, R.F. Lobo, Chem. Mater. 4 Ž1992. 756. Chem. Commun. Ž1991. 57.
w7x Y. Kubota, M.M. Helmkamp, S.I. Zones, M.A. Davis, Mi- w37x D. Chatterjee, H.C. Bajaj, A. Das, K. Bhatt, J. Catal. 92
croporous Mater. 6 Ž1996. 213. Ž1994. 235.
w8x S. Feast, D. Bethell, P.C. Bulman Page, F. King, C.H. w38x A. Corma, M. Iglesias, C. del Pino, F. Sanchez,
´ J. Chem.
Rochester, M.R.H. Siddiqui, D.J. Willock, G.J. Hutchings, J. Soc., Chem. Commun. Ž1991. 1253.
Chem. Soc., Chem. Commun. Ž1995. 2409. w39x A. Corma, M. Iglesias, C. del Pino, F. Sanchez,
´ Stud. Surf.
w9x S. Feast, D. Bethell, P.C. Bulman Page, F. King, C.H. Sci. Catal. C 75 Ž1993. 2293.
Rochester, M.R.H. Siddiqui, D.J. Willock, G.J. Hutchings, J. w40x W. Reschetilowski, U. Bohmer,
¨ J. Wiehl, Stud. Surf. Sci.
Mol. Catal. A 107 Ž1996. 291. Catal. C 84 Ž1994. 2021.
w10x P.B. Weisz, V.J. Frilette, J. Phys. Chem. 64 Ž1960. 382. w41x U. Bohmer,
¨ K. Morgenschweis, W. Reschetilowski, Catal.
w11x V.J. Frilette, N.J. Erlton, P.B. Weisz, US Pat., 3.140.322, Today 24 Ž1995. 195.
1964. w42x W. Reschetilowski, U. Bohmer,
¨ K. Morgenschweis, Chem.
w12x P.B. Weisz, R.W. Maatman, E.B. Mower, J. Catal. 1 Ž1962. Ing. Tech. 67 Ž1995. 205.
307. w43x N. van der Puil, E.C. Rodenburg, H. van Bekkum, J.C.
w13x P.B. Weisz, Chemtech 3 Ž1973. 498. Jansen, Stud. Surf. Sci. Catal. 97 Ž1995. 377.
w14x N.Y. Chen, P.B. Weisz, Chem. Eng. Progr. Symp. Ser. 73 w44x N. van der Puil, E.J. Creyghton, E.C. Rodenburg, T.S. Sie,
Ž1967. 86. H. van Bekkum, J.C. Jansen, J. Chem. Soc., Faraday Trans.
w15x N.Y. Chen, U.S. Pat., 3.496.246, 1970. 92 Ž1996. 4609.
w16x F. Steinbach, H. Minchev, Z. Phys. Chem. 99 Ž1976. 235. w45x J. Shabtai, R. Lazar, E. Biron, J. Catal. 27 Ž1984. 35.
w17x R.M. Dessau, J. Catal. 77 Ž1982. 304. w46x E.J. Creyghton, S.D. Ganeshie, R.S. Downing, H. van
w18x R.M. Dessau, J. Catal. 89 Ž1984. 520. Bekkum, J. Chem. Soc., Chem. Commun. Ž1995. 1859.
w19x R.M. Dessau, US Pat., 4.377.503, 1983; Mobil Oil, USA, JP w47x E.J. Creyghton, S.D. Ganeshie, R.S. Downing, H. van
Pat., 59.156.437, 1983. Bekkum, J. Mol. Catal. A 115 Ž1996. 457.
w20x J. Weitkamp, T. Kromminga, S. Ernst, Chem. Ing. Tech. 64 w48x J.C. van der Waal, K. Tan, H. van Bekkum, Catal. Lett. 41
Ž1992. 1112. Ž1996. 63.
w21x J. Weitkamp, S. Ernst, T. Bock, A. Kiss, P. Kleinschmit, w49x J.C. van der Waal, P. Lin, M.S. Rigutto, H. van Bekkum,
Stud. Surf. Sci. Catal. 94 Ž1995. 278. Stud. Surf. Sci. Catal. 105 Ž1997. 1093.

Vous aimerez peut-être aussi