Sie sind auf Seite 1von 10

Microporous and Mesoporous Materials 35–36 (2000) 21–30

www.elsevier.nl/locate/micromeso

Current views on the mechanism of catalytic cracking


A. Corma a, *, A.V. Orchillés b
a Instituto de Tecnologı́a Quı́mica, UPV-CSIC, Avenida de los Naranjos s/n, 46022 Valencia, Spain
b Departament d’Enginyeria Quı́mica, Universitat de València, Dr. Moliner, 50, 46100 Burjassot, Spain

Received 25 March 1999; received in revised form 12 July 1999; accepted for publication 13 July 1999

Dedicated to the late Werner O. Haag in appreciation of his outstanding contributions to heterogeneous catalysis and zeolite science

Abstract

The cracking mechanisms of hydrocarbons have been reviewed and the kinetic and thermodynamic implications
of the different steps, i.e. initiation, chain propagation, and termination, have been discussed. Although the cracking
mechanism of olefins and alkylaromatics is well established, the initiation step for the cracking of paraffins is still
under debate. The role of Brönsted-type active sites and also the possible influence of extra-framework Al species in
the case of zeolite catalysts, especially when commercial feeds and industrial conditions are employed, are presented.
The product distribution is determined by the number of propagation events occurring per initiation step, and this
can be controlled by carefully selecting the precise reaction conditions and the catalyst (including its composition and
pore topology). © 2000 Elsevier Science B.V. All rights reserved.

Keywords: Carbocations; Catalytic cracking of hydrocarbons; Cracking chain mechanism; Cracking mechanism; Paraffin activation

1. Introduction these areas that the work of Werner Haag has


been most important:
It is fair to say that the catalytic cracking of $ What is the exact nature of the active sites?
hydrocarbons ( FCC ) is one of the most important $ What is the nature of the transition state?
processes in the oil refining industry. Owing to $ Is there only one route or are there more routes
this, and to the existence of detailed mechanistic through which the reaction proceeds?
proposals that allow us to rationalize the results In this paper we will try to summarize the current
obtained, it has been an active research field, with state of knowledge in this important field.
more than 5000 articles published during the last
10 years alone.
However, despite the considerable advance- 2. Catalytic cracking mechanism: the initiation step
ments in this area and the enormous amount of
work performed, there are still three general areas The catalytic cracking of hydrocarbons is a
where we still require more understanding. It is in chain reaction that is believed to follow the carbo-
nium ion theory developed by Whitmore [1]. This
chain mechanism involves three elementary steps:
* Corresponding author. Tel.: +34-96-387-7800; initiation, propagation and termination. The initia-
fax: +34-96-387-7809. tion step is represented by the attack of an active
E-mail address: itq@upvnet.upv.es (A. Corma) site on the reactant molecule to produce the acti-

1387-1811/00/$ - see front matter © 2000 Elsevier Science B.V. All rights reserved.
PII: S1 3 8 7 -1 8 1 1 ( 9 9 ) 0 0 20 5 - X
22 A. Corma, A.V. Orchillés / Microporous and Mesoporous Materials 35–36 (2000) 21–30

vated complex that, in the gas phase or when using


liquid superacids, would correspond to the forma-
tion of a carbocation [2]. The chain propagation
is represented by the transfer of a hydride ion
from a reactant molecule to an adsorbed carbe-
nium ion. Finally, the termination step corres-
ponds to desorption of the adsorbed carbenium
ion to give an olefin whilst restoring the initial
active site.
In more detail, the initiation step, if one con-
siders an olefin as the reactant, corresponds to the
attack of one Brönsted acid site from the catalyst
to the double bond of the reactant olefin to form
a carbenium ion. This is certainly what occurs in
the gas phase [3], or in the presence of liquid Fig. 1. A concerted mechanism for charge isomerization [7].
superacids where the positive charge is stabilized
by the anion [4].
When this carbocation is formed it can
rearrange without increasing the branching of the
chain or can rearrange to generate a branched
carbocation. The former process is believed to
occur through a 1,2 hydrogen shift, whereas
branching isomerization occurs via a protonated
cyclopropane ring as the intermediate [5].
When the olefin interacts with solid acids, and,
more specifically, with zeolites, weak Brönsted acid
sites are already active and the reactions described
above can readily occur. Kazansky [6 ] proposed
that the activation of the reactant and the isomer-
ization process occurred in two steps. Firstly the
formation of a surface alkoxide by addition of a
proton from the zeolite, and in a second step the Fig. 2. Transition state of branching isomerization in linear
alkoxide decomposes to give the isomerized pro- butenes over zeolites. Distances in angstroms and angles in
duct. However, a mechanism like the one proposed degrees [7].
above involving highly stable alkoxide species also
involves activation energies much higher than be shown later, this can play an important role
those found experimentally. This problem has been during the cracking of linear paraffins and olefins.
recently revisited [7] and it has been shown that, A third reaction, which can occur when a proton
after protonation of the olefin, the charged isomer- has attacked an olefin, is the cracking of the CMC
ization process can occur via a concerted mecha- bond located in b position with respect to the
nism through the transition state given in Fig. 1. carbon supporting the positive charge (b scission)
The formation of this transition state involves [10,11]. During this process the formation of pri-
16.1 kcal mol−1; this is in good agreement with mary carbenium ions should be avoided, and this
the observed experimental values, which range occurs if 1,2 hydrogen shifts and branching isomer-
between 15 and 20 kcal mol−1 [8,9]. Branching ization take place prior to the CMC cracking
isomerization also occurs on zeolites, via a proton- process. For carbenium ions with a carbon number
ated cyclopropane ring (PCP) (Fig. 2) and, as will less than six, no rearrangements are possible that
A. Corma, A.V. Orchillés / Microporous and Mesoporous Materials 35–36 (2000) 21–30 23

can avoid the formation of primary carbenium by the direct attack of a Brönsted acid site on a
ions at some point of process cracking, whereas CMC bond, forming a surface ‘carbonium ion
octene cracking can proceed via tertiary carbenium type’ activated complex [19,20]. When this carbo-
ion intermediates. This explains the dramatic nium ion is formed it can crack via protolytic
decrease with carbon number of the cracking rates cracking or even give hydrogen if the proton
of C5–C8 olefins, as indicated by Buchanan et al. attack takes place at a CMH bond. Using
[12]. 3-methylpentane and n-hexane as reactants, Haag
A general agreement exists in the literature on and Dessau [18] proposed the reaction scheme and
the nature of the active sites and the global mecha- the active sites for the initiation step in the cracking
nism through which the cracking of olefins and of paraffins shown in Fig. 3.
alkylaromatics occur. However, this is not the case It can be seen there that the protonic attack
for the cracking of paraffins, where the initiation occurs on the most nucleophilic centres of the
step has been widely debated. It was first thought molecule, i.e. on the s CMC and CMH bonds
that the presence of olefins, even at trace levels, [21]. The protonation of the CMH bonds does
was necessary to initiate the cracking of a paraffin not occur on the primary carbon atoms, as is
[13,14]. In this mechanism the traces of olefins are demonstrated by the results from neopentane
readily protonated and the carbenium ions formed cracking on Y zeolites [22], and H is mainly
2
are able to abstract a hydride ion from a paraffin, obtained from paraffins containing tertiary carbon
generating in this way the carbenium ion necessary atoms [23,24]. On the other hand, the formation
for the paraffin to keep reacting: of carbonium ions and their rupture only obeys
statistical criteria and does not involve large ener-
+H+ C getic barriers. In this way the rates of formation
R–CHNCH–R∞  R–CH–CH –R∞ of H , CH and C H during cracking of n-butane
2 2 4 2 6
C are practically the same [25], and when the mecha-
R–CH−CH –R∞+CH –CH –CH –CH –CH  nism is studied by means of molecular orbital
2 3 2 2 2 3
C calculations [26 ] it is found that the lower activa-
R–CH –CH –R∞+CH –CH –CH–CH –CH –CH . tion energy required for the protolytic cracking at
2 2 3 2 2 2 3
the central atom is compensated by the larger
It was also claimed that the olefins necessary to
activation entropy of the external bond. This com-
initiate the cracking of paraffins could be formed pensation makes equally probable the rupture of
by the direct attack of a proton on an HMC bond any of the CMC bonds in n-butane.
of the paraffin to give H and an olefin in a similar After the introduction of protolytic cracking as
2
way as occurs in the presence of liquid super- the initiation step it appeared that the mechanistic
acids [15]. problems for the cracking of paraffins were solved.
The formation of the first surface carbenium However, several points have been raised recently
ion was also proposed to occur by abstraction of that suggest further discussion is still needed to
a hydride ion by a Lewis acid site of the catalyst resolve this matter fully. Firstly, Sommer and
[16,17]. In all cases, regardless of the active site coworkers [27–29] have produced an excellent
involved, the authors always proposed that the piece of work on the initiation and chain propaga-
cracking of paraffins on solid acids occurred tion steps during the reaction of isobutane with
through the formation of a carbenium ion as the deuterated liquid superacids using deuterated
initiation step. It is surprising that the authors did solids at temperatures in the range of 20–200°C.
not take into consideration the paraffin cracking Their findings show that the H/D exchange
mechanism, which was demonstrated to work in between the liquid superacid and isobutane takes
liquid superacids, and which showed that the initia- place in the primary as well as in the tertiary CMH
tion could also occur by direct protonation of a groups. The observed regioselectivity is in
CMC bond [4]. This remained so until 1984, when agreement with the relative basicity of the s CMH
Haag and Dessau [18] proposed that also on solid bonds. In this case a reaction scheme via the
acids the cracking of paraffins could be initiated formation of pentacoordinated ions can explain
24 A. Corma, A.V. Orchillés / Microporous and Mesoporous Materials 35–36 (2000) 21–30

Fig. 3. Protolytic cracking of 3-methylpentane and n-hexane via pentacoordinated carbonium ions [18].

the mechanism of activation and isotopic exchange becomes protonated (see Fig. 5). In other words,
(Fig. 4). the formation of an olefin will be the first step
On the contrary, with a solid acid catalyst, or instead of the direct attack of the proton on a
even with solid superacids, the H/D exchange only CMC bond. Finally, the same authors conclude
occurs on the primary hydrogen of the isobutane. that, at lower temperatures on solid acids, the
Then, this behaviour can only be explained if one activation mechanism of paraffins has not yet been
assumes that isobutene is formed firstly and this clearly demonstrated without ambiguities.

Fig. 4. Exchange and protolysis equilibria of isobutane in DF–SbF [27–29].


5
A. Corma, A.V. Orchillés / Microporous and Mesoporous Materials 35–36 (2000) 21–30 25

Fig. 6. Proposed mechanism of alkane cracking [33,34].


Fig. 5. Catalytic cycle of H/D exchange on solid acids [AH ]
[27–29].
bond in the a position with respect to the active
sites, which is always located on the tertiary carbon
Moreover, it has been suggested [15–17] that ( Fig. 6). This mechanism predicts large amounts
the presence of Lewis acid or electron acceptor of olefins, which implies the formation of molecu-
sites in many solid acids (and certainly in zeolites lar H . Although this is an interesting proposal, it
2
where they are associated with partially coordi- is unfortunate that the working temperature used
nated framework Al, or extra-framework Al by the author was far too low (150°C ) and,
( EFAL) species [30]) can enhance the dehydroge- furthermore, the amount of H produced was not
2
nation of paraffins, thus forming olefins that can determined experimentally.
initiate, or at least enhance, the cracking reaction. In conclusion, current knowledge certainly sug-
In isobutane cracking, this possibility was firstly gests that Brönsted active sites intervene in the
postulated by McVicker et al. [31], who proposed initiation step for the cracking of olefins and
an initiation step based in the generation of a alkylaromatics, and also in the possible protolytic
radical cation over zeolite electron acceptor sites. cracking of paraffins. However, the generation of
In this way, isobutene formation was assumed to olefins from paraffins by Brönsted and Lewis acid
occur by decomposition of a surface-bound isobu- sites has also to be considered as an initiation or,
tane radical cation. This hypothesis could have a at least, as a complementary initiation mechanism
renewed interest with the work done by the groups during the cracking of paraffins.
of Sommer and Dumesic [32]. We will return to If one now tries to use this accumulated knowl-
this point later. edge on the cracking mechanism generated from
Recently, Kissin [33,34] has proposed a new more fundamental studies using pure compounds
paraffin cracking initiation mechanism in which to explain what is occurring during the cracking
Brönsted acid sites are involved. The author firstly of commercial feeds, one must also consider the
noted a different product distribution when crack- following. In commercial units the temperature at
ing olefins and isoalkanes with the same structure. which the feed encounters the catalyst is quite high
Thus, although the product distribution obtained (~700°C ); moreover, the zeolite is not a ‘pure’
during cracking of olefins is consistent with a b Brönsted acid catalyst. In addition, a large amount
scission mechanism, this is not the case for the of dispersed EFAL also exists on the surface of
isoparaffins. More specifically, a larger amount of the zeolite crystals. This EFAL presents Lewis
short chain products are produced during the acidity, and will be the first type of site that will
cracking of isoalkanes. To explain this, the author encounter the large reactant molecules present in
proposes that the cracking occurs at the CMC the feed. It can thus be expected that, in accordance
26 A. Corma, A.V. Orchillés / Microporous and Mesoporous Materials 35–36 (2000) 21–30

with the opinion of McVicker et al. [31] and


Corma and coworkers [30,37], besides protolytic
cracking, a ‘radical’ type of cracking and dehydro-
genations will also occur on this EFAL and olefins
will be generated, even if one starts from a purely
paraffinic feed. These olefins can, at the very least,
enhance the rate of paraffin cracking by adsorbing
on Brönsted acid sites and form a ‘carbenia sur-
face’ [35], which can activate paraffins by hydride
abstraction just as Lewis acid sites would do.
Good support for this idea comes from the work
of Schuette and Schweizer [36 ], who enhanced the
cracking rate of industrial feeds by introducing
Lewis acidity via the introduction of active
alumina. Fig. 7. Hydride transfer between different carbenium ions and
paraffins in the gas phase: (1) (C H MHMC H )+; (2)
Therefore, if we wish to understand better what 2 5 2 5
(C H MHMC H )+; (3) (tert-C H MHMtert-C H )+; (4)
is occurring during the cracking in industrial units, 3 7 3 7 4 9 4 9
(C H MHMtert-C H )+ [39].
further research under more realistic experimental 3 7 4 9
conditions and feeds is required. This has been
attempted by carrying out the cracking of n- isomerize and/or crack, therefore keeping the reac-
heptane and vacuum gasoil on USY zeolites with tion chain going. When this interaction was care-
a similar unit cell size but with different contents fully studied in the gas phase [39], it was found
of EFAL [37]. It has been observed that when that a stable carbocation intermediate is formed
EFAL is removed, while preserving the unit cell ( Fig. 7). This presents a CMHMC bond that is
size of the USY zeolite, gasoil conversion is bielectronic and tricentric, and whose geometry
decreased and lower selectivities to liquid distillates changes depending on the carbon atoms
are obtained. These experiments clearly showed implicated.
that, at least for gasoil cracking, the Lewis acidity Similar results have been obtained when the
associated with EFAL is also intervening in the hydride transfer process occurs between a paraffin
reaction. and a carbenium ion adsorbed on a zeolite cluster.
At this point it should be possible to write a
general mechanism for paraffin cracking [18,41] in
3. Cracking mechanism: chain propagation which the initiation step occurs through a mono-
molecular reaction where the ‘carbocation’ is
Regardless of the initiation and cracking mecha- formed. This will be followed by a bimolecular
nism, when the CMC bond is broken a ‘carbenium- chain transfer reaction that corresponds to the
like ion’ is left on the surface. This can either hydride transfer between an adsorbed ‘carbenium
desorb, giving one olefin or one paraffin when the ion’ and a reactant molecule:
cracking occurs via protolytic or b-scission respec- After the hydride transfer has occurred and a
tively, or can interact with one molecule of reac- ‘carbenium ion’ of the reactant molecule is formed,
tant. For the case of a paraffin the interaction with it sometimes becomes difficult to visualize a
a carbenium ion has been studied theoretically b-scission without involving a primary carbenium
[38–40], and it has been traditionally proposed ion. Thus, owing to the high instability of primary
that chain propagation will occur when a carbe- carbenium ions, a search has been made for alter-
nium ion abstracts a hydride ion from a reactant native cracking routes that avoid the formation of
paraffin, and while the former will desorb as a primary carbocations. In this way Sie [42,43] has
paraffin, the reactant molecule will be converted proposed a new mechanism for the cracking of
into a carbenium ion that will be able to either carbenium ions that involves the participation of
A. Corma, A.V. Orchillés / Microporous and Mesoporous Materials 35–36 (2000) 21–30 27

PCP structures [44]. The mechanism is presented obtained if only product distribution data are used
in Fig. 8. in the procedure [45]. Nevertheless, one may use
This mechanism not only avoids the participa- isobutane as a reactant molecule to test the ability
tion of primary carbenium ions, but also can of different catalysts to carry out both monomolec-
explain the high yield of isoparaffins obtained ular protolytic cracking and bimolecular cracking.
during the cracking of n-paraffins. Isobutane, in addition to the tertiary carbon, has
There is no doubt that the product distribution a tertiary CMH bond that can easily react by
observed during cracking of paraffins will depend either a proton attack or by a hydride transfer
on the relative contribution of protolytic and reaction [24].
b-scission mechanisms of CMC rupture. Then, if As a consequence, isobutane gives H , CH , n-
2 4
most of the cracking occurs through a bimolecular butane, propane and i-butene as the primary pro-
b-scission mechanism, high yields of branched ducts. Among them, only H and CH are gener-
2 4
products will be obtained. On the contrary, if ated by the protolytic route, whereas the rest of
protolytic cracking dominates, then more lineal the products are obtained by the bimolecular path-
paraffins, methane, ethane and even ethylene and way involving hydride transfer and b-scission. It
H will be produced. is then easy, in this case, not only to find the
2
Based only on the product distribution relative contribution of each of those two mecha-
obtained, some authors [16,17] have made a pro- nisms on a given catalyst [23,24,46 ], but also to
posal to evaluate the contribution of both types calculate a ‘chain length’ (number of chains per
of cracking mechanism taking place on a given initiating event) by means of the ratio between the
catalyst and with a given reactant. However, a bimolecular propagation and the monomolecular
deep consideration of all the reactions occurring reaction [24]. It is obvious that the ‘chain length’
during the cracking of paraffins indicates that, will depend on the average lifetime of the carbe-
when reactants with more than four carbon atoms nium ion or its concentration on the catalyst
are used, it becomes impossible to determine the surface, and this in turn will be a function of the
exact contribution of the protolytic and b-scission nature of the reactant and the catalyst. More
mechanism to the final product distribution specifically, in the case of zeolites there is an
28 A. Corma, A.V. Orchillés / Microporous and Mesoporous Materials 35–36 (2000) 21–30

[24,41]. This step is just the reverse of the adsorp-


tion of one olefin and could be written as:
C H+C H +H+.
3 7 3 6
With this simple step, the cracking chain mecha-
nism is completed.

5. Influence of the catalyst and operation variables


on the reaction mechanism

Haag and Dessau [18] found that the activation


energy was higher for the protolytic than for the
b-scission mechanism. Thus, it can be expected
that the monomolecular mechanism will predomi-
nate at high reaction temperatures. Moreover,
the monomolecular character of the protolytic and
the bimolecular nature of b-scission will make the
former predominant at low partial pressures of
hydrocarbon and lower levels of conversion. These
conclusions were supported by the work of Riekert
and Zhou [49], who observed an induction period
during paraffins cracking (which is typical of bimo-
lecular reactions) only at reaction temperatures
below 200°C. In an analogous way, Lukyanov
et al. [50] have only observed an important contri-
bution of the hydride transfer on the cracking of
n-hexane with ZSM-5 at reaction temperatures
below 400°C.
Furthermore, since the bimolecular hydride
Fig. 8. Proposed mechanism of acid-catalysed cracking of transfer leading to a b-scission mechanism involves
normal paraffins [42]. a larger transition state than the unimolecular
protolytic cracking, it appears that the pore size
of the zeolite used as the catalyst can introduce a
increase in the ‘chain length’ when the framework shape selectivity effect that will control the relative
Si/Al ratio is increased [23,24]. On the other hand, extension of the protolytic and b-scission mecha-
detailed kinetic studies, including connecting cata- nisms. On the other hand, we must take into
lytic cycles, have also allowed the separation of account that bimolecular mechanisms are favoured
protolytic cracking from the other reactions. In at higher site density and higher adsorption capac-
this way the rate kinetic constants for the different ity. Thus, an increase of the framework Si/Al ratio
groups of reactions occurring during catalytic of zeolite will have a negative effect on the rate of
cracking have been able to be calculated [47,48]. bimolecular cracking. All of this will certainly be
reflected by the product distribution obtained.
Giannetto et al. [51] concluded that bimolecular
4. Cracking mechanism: chain termination cracking is more restricted in Y zeolites with a
high Si/Al ratio. Haag and Dessau [18] found that
The chain transfer is terminated when the sur- protolytic cracking is more favourable in a 10MR
face ‘carbenium ions’ are desorbed and the zeolite, such as ZSM-5, than in a 12MR zeolite,
Brönsted acid site of the catalyst is regenerated such as FAU. Mirodatos and Barthomeuf [52]
A. Corma, A.V. Orchillés / Microporous and Mesoporous Materials 35–36 (2000) 21–30 29

Fig. 10. Protolytic/hydride transfer ratio versus pore dimen-


sions in n-heptane cracking over different zeolites. (&) USY;
(+) ZSM-5; (×) Mordenite; (#) Beta; ($) MCM-22; (%)
CIT-1; (6) SSZ-24; (+) NU-86; ()) NU-87; (1) Ferrierite;
(n) ITQ-4.

olefin/paraffin production. Moreover, it will pro-


duce less branched products and lower selectivity
Fig. 9. Variation of cracking selectivity as a function of zeolite to gasoline and diesel.
structure [53]. On the other hand, the bimolecular mechanism
will give less dry gas, more gasoline and diesel,
further supported these findings and proposed that lower olefins, and probably more catalytic coke.
protolytic cracking is favoured when small-pore Thus, catalyst manufacturers can utilize this
zeolites and/or zeolites with highly tortuous pores knowledge and modify the zeolite Y present in the
are used. Wielers et al. [53] showed that the catalyst by stabilizing it with a higher or lower
cracking mechanism ratio (CMR), defined as the framework Si/Al ratio depending on whether a
ratio C +C /isobutane (which is directly propor- high yield of olefins and LPG olefins is desired
1 2
tional to the ratio of protolytic to b-scission mecha- instead of a higher yield of distillates with a low
nism), increased with decreasing zeolite pore yield of gases respectively. Moreover, it is well
diameter during n-hexane cracking ( Fig. 9). In an known that the introduction of ZSM-5 as a zeolite
analogous way, when a protolytic versus hydride additive should, in addition, contribute to increase
transfer ratio was obtained on the basis of the the ratio of protolytic versus b-scission cracking,
paraffins and olefins produced during cracking of therefore augmenting the olefinicity of LPG and
n-heptane [54], it was also found that this ratio gasoline, but decreasing the gasoline yield.
increases when decreasing the pore dimensions of
the zeolite ( Fig. 10).
The ratio of the protolytic versus the b-scission 6. Conclusions
mechanism is important not only from a funda-
mental standpoint, but has important implications It has been discussed that although the cracking
for a real FCC unit. Indeed, the product distribu- mechanism of olefins and alkylaromatics is well
tion obtained by a purely protolytic cracking will established, the activation step for paraffins still
be very similar to that obtained by a radical- needs further work to reach the required level of
type mechanism, i.e. larger H , CH , C and understanding. From our current knowledge it
2 4 2
30 A. Corma, A.V. Orchillés / Microporous and Mesoporous Materials 35–36 (2000) 21–30

appears that the most probable activation route, [22] E.A. Lombardo, R. Pierantozzi, K. Hall, J. Catal. 110
at least at high reaction temperatures, involves the (1988) 171.
[23] A. Corma, P.J. Miguel, A.V. Orchillés, J. Catal. 145
attack of a proton from the catalyst on a CMC (1994) 171.
bond or on a CMH bond, especially if there are [24] P.V. Shertudke, G. Marcelin, G.A. Sill, W.K. Hall,
tertiary hydrogen atoms present. However, it has J. Catal. 36 (1992) 446.
been shown that the role of EFAL species, particu- [25] H. Krannila, W.O. Haag, B.C. Gates, J. Catal. 135
larly when working with commercial feeds and (1992) 115.
[26 ] S.J. Collins, P.J. O’Malley, J. Catal. 153 (1995) 94.
industrial conditions, cannot be neglected. [27] J. Sommer, A. Sassi, M. Hachoumy, R. Jost, A. Karlsson,
Finally, the product distribution observed is P. Ahlberg, J. Catal. 171 (1997) 391.
highly dependent on the chain propagation mecha- [28] J. Sommer, R. Jost, M. Hachoumy, Catal. Today 38
nism, and more specifically on the number of (1997) 309.
[29] J. Sommer, D. Habermacher, M. Hachomy, R. Jost, A.
propagation events per initiation step. This param-
Reynaud, Appl. Catal. A: Gen. 146 (1996) 193.
eter can be controlled by the reaction conditions, [30] A. Corma, V. Fornés, A. Martı́nez, A.V. Orchillés, ACS
as well as by the nature (composition and topol- Symp. Ser. 368 (1988) 542.
ogy) of the catalyst-zeolite used. [31] G.B. McVicker, G.M. Kramer, J.J. Ziemiak, J. Catal. 83
(1983) 286.
[32] Z. Hong, K.B. Fogash, R.M. Watwe, B. Kim, B.I.
Masqueda-Jimenez, M.A. Natal-Santiago, J.M. Hill, J.A.
References Dumesic, J. Catal. 178 (1998) 489.
[33] Y.V. Kissin, J. Catal. 163 (1996) 50.
[1] F.C. Whitmore, Ind. Eng. Chem. 26 (1934) 94. [34] Y.V. Kissin, J. Catal. 180 (1998) 101.
[2] V.B. Kazansky, M.V. Frash, R.A. Van Santen, Appl. [35] B.W. Wojciechowski, A. Corma, Catalytic Cracking. Cata-
Catal. A.: Gen. 146 (1996) 225. lysts, Chemistry and Kinetics, Marcel Dekker, 1986.
[3] K. Hiraoka, P. Kebarle, J. Am. Chem. Soc. 98 (1976) 6119. [36 ] L.W. Schuette, A.E. Schweizer, EP 749 781, 1996.
[4] G.A. Olah, J. Am. Chem. Soc. 95 (1973) 4939. [37] A. Corma, Stud. Surf. Sci. Catal. 49 (1989) 49.
[5] D.M. Brower, in: R. Prins, G.C.A. Schmit ( Eds.), Chemis- [38] M.V. Frash, V.N. Solkan, V.B. Kazansky, J. Chem. Soc.
try and Chemical Engineering of Catalytic Processes, Faraday Trans. 93 (1997) 515.
Sijthoff and Noordhoff, Alphen aan den Rijn, The Nether- [39] M. Boronat, P. Viruela, A. Corma, J. Phys. Chem. A 101
lands, 1980, p. 137. (1997) 10 069.
[6 ] V.B. Kazansky, Stud. Surf. Sci. Catal. 84 (1994) 251. [40] M. Boronat, P. Viruela, A. Corma, J. Phys. Chem. A 102
[7] M. Boronat, P. Viruela, A. Corma, J. Phys. Chem. A 102 (1998) 9863.
(1998) 982. [41] A. Corma, J.B. Montón, A.V. Orchillés, Appl. Catal. 23
[8] V.B. Kazansky, Acc. Chem. Res. 24 (1991) 379. (1986) 255.
[9] A. Corma, Chem. Rev. 95 (1995) 559. [42] S.T. Sie, Ind. Eng. Chem. Res. 31 (1992) 1881.
[10] C.L. Thomas, Ind. Eng. Chem. 41 (1949) 2564. [43] S.T. Sie, Ind. Eng. Chem. Res. 32 (1993) 397.
[11] B.S. Greensfelder, H.H. Voge, Ind. Eng. Chem. 37 [44] J. Weitkamp, Ind. Eng. Chem. Prod. Res. Dev. 21
(1945) 983. (1982) 550.
[12] J.S. Buchanan, J.G. Santiesteban, W.O. Haag, J. Catal. [45] A. Corma, P.J. Miguel, A.V. Orchillés, J. Catal. 172
158 (1996) 279. (1997) 355.
[13] B.S. Greensfelder, H.H. Voge, G.M. Good, Ind. Eng. [46 ] Y. Ono, K. Kanae, J. Chem. Soc. Faraday Trans. 87
Chem. 41 (1949) 2573. (1991) 663.
[14] W.F. Pansing, J. Phys. Chem. 69 (1966) 392. [47] G. Yaluris, J.E. Rekoske, L.M. Aparicio, R.J. Madon, J.A.
[15] T.F. Narbeshuber, A. Brait, K. Seshan, J.A. Lercher, Appl. Dumesic, J. Catal. 153 (1995) 54.
Catal. A: Gen. 146 (1996) 119. [48] R.A. Watson, M.T. Klein, R.H. Harding, Ind. Eng. Chem.
[16 ] A. Brait, A. Koopmans, H. Weinstabe, A. Ecker, K. Res. 35 (1996) 1506.
Seshan, J.A. Lercher, Ind. Eng. Chem. Res. 37 (1998) 873. [49] L. Riekert, J. Zhou, J. Catal. 137 (1992) 437.
[17] S.E. Tung, E.J. McIninch, J. Catal. 10 (1968) 166. [50] D.B. Lukyanov, V.I. Shtral, S.N. Khadzhiev, J. Catal. 146
[18] W.O. Haag, R.M. Dessau, Proceedings of the 8th Interna- (1994) 87.
tional Congress on Catalysis, Dechema, Berlin 2 (1984) 305. [51] G. Giannetto, S. Sansare, M. Guisnet, J. Chem. Soc.
[19] A. Corma, J. Planelles, J. Sánchez-Marı́n, F. Tomás, Chem. Commun. (1986) 1303.
J. Catal. 93 (1985) 30. [52] C. Mirodatos, D. Barthomeuf, J. Catal. 114 (1989) 121.
[20] C.J.A. Mota, P.M. Esteves, A. Ramirez-Solis, R. Hernan- [53] A.F.H. Wielers, M. Vaarkamp, M.F.M. Post, J. Catal. 127
dez-Lamoneda, J. Am. Chem. Soc. 119 (1997) 5193. (1991) 51.
[21] S.R. Blaszkowski, M.A.C. Nascimento, R.A. Van Santen, [54] A. Corma, V. González-Alfaro, A.V. Orchillés, Appl.
J. Phys. Chem. 100 (1996) 3463. Catal. A: Gen. (1999) in press.

Das könnte Ihnen auch gefallen