Anda di halaman 1dari 36

Microporous and Mesoporous Materials 82 (2005) 257292

www.elsevier.com/locate/micromeso

Gas phase catalysis by zeolites


Michael Stocker

SINTEF Materials and Chemistry, Department of Hydrocarbon Process Chemistry, P.O. Box 124 Blindern, N-0314 Oslo, Norway
Received 25 April 2004; received in revised form 15 November 2004; accepted 20 November 2004
Available online 8 April 2005
Dedicated to Diplom-Chemiker Ulf Blindheim on the occasion of his 70th birthday

Abstract
This paper provides an overview about todays use of zeolites and related microporous materials as catalysts within the elds of
rening, petrochemistry and commodity chemicals. The content of this presentation is devoted to gas phase catalysiswith focus on
acid catalysis, hydrocarbon conversion and formation, oil and natural gas upgrading as well as catalytic probe reactions for the
characterisation of zeolites and related microporous materials. The review is primarily meant for beginners who intend to get
acquainted with this eld. However, for more detailed information the interested reader is invited to consult the dedicated papers
cited throughout this overview.
2005 Elsevier Inc. All rights reserved.
Keywords: Zeolites; Microporous materials; Gas phase catalysis; Crude oil upgrading; Natural gas conversion

1. Introduction
Catalysis by zeoliteswith focus on hydrocarbon
conversion and formationcovers nowadays a broad
range of processes related to the upgrading of crude oil
and natural gas. This includes, among others, uid catalytic cracking (FCC), hydrocracking, dewaxing, aliphate
alkylation, isomerisation, oligomerisation, transformation of aromatics, transalkylation, hydrodecyclisation
as well as the conversion of methanol to hydrocarbons.
All these conversions are catalysed by zeolites or related
microporous materials, based both on the acid properties and shape-selective behaviour of this type of
materials.
The rst part of this chapter deals with the understanding of the chemistry of acid catalysis using zeolites
or related microporous materials, including the formation of acid sites, carbocation chemistry and their
*

Tel.: +47 98 24 39 33; fax: +47 22 06 73 50.


E-mail address: michael.stocker@sintef.no

1387-1811/$ - see front matter 2005 Elsevier Inc. All rights reserved.
doi:10.1016/j.micromeso.2005.01.039

reaction mechanisms, as well as the importance of the


shape-selectivity of the microporous materials. The term
zeolite is used for the microporous aluminosilicate
systems, however, SAPO type catalysts belong to the
family of microporous materials as well.
The second part of this chapter covers the discussion
of the present situation and the new developments related to the above mentioned petroleum rening and
natural gas conversion processes using microporous
materials, with focus on the new requirements due to
the introduction of new fuel specications world-wide.
Finally, the third part of this chapter is dedicated to
the dierent probe reactions with respect to the characterisation of zeolites and related microporous materials.
Since the micropores of zeolites and related compounds
have diameters in the range of molecular dimensions,
the shape-selective eect reveals unique possibilities in
the catalytic conversion of, for example, hydrocarbons.
However, proper utilisation of this behaviour requires
that the conversion occurs at active sites connected
to the internal pore structure and not at the external

258

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

surface of the zeolite crystals. A suitable possibility for


the investigation of relative activities of the internal
and external surface of microporous solids is the application of appropriate probe molecules, either for the
investigation of their interaction with the internal and
external surface of the zeolite or as model compounds
during catalytic conversions.

2. Acid catalysis and shape selectivity of zeolites and


related microporous materials with respect to
hydrocarbon conversion and formation

Fig. 1. Brnsted acid sites (bridging hydroxyl groups) in zeolites


[1,2].

2.1. Acid sites


Zeolites and related microporous molecular sieves
consist of a three-dimensional network of metaloxygen
tetrahedra (in a few cases also octahedra) which provide
the periodically sized microporous structure, in which
the active sites are part of the structure. Acid sites result
from the imbalance of the metal and the oxygen formal
charge in the primary building unit. This can easily be
recognised in the case of zeolites, which consist of a
three-dimensional network of SiO tetrahedra. A lattice
comprising of only SiO tetrahedra is neutral (the 4+
charge at the silicon is balanced by four oxygen atoms
with each 2 charge, however, belonging to two tetrahedra). Replacing one Si4+ atom by Al3+ causes a formal
charge on the tetrahedron of 1. This negative charge
is then balanced by a proton or metal cation forming
an acid site. The bare, negatively charged tetrahedron
is then the corresponding base. Please keep in mind that
these acid and base properties are not just a function of
the chemical composition, since other factors, like the
framework density, the type of cation or the local strain
have an inuence as well [1].
In AlPO4 type microporous materials the framework
structure consists of a strictly alternating AlOP sequence (Al3+ and P5+, balanced by four oxygen atoms
with each 2 charge, however, belonging to two tetrahedra), resulting in a completely neutral lattice as well, like
in the case of pure silica zeolites. Depending on the combinations of the metal cation in the lattice, frameworks
with positive or negative charges are in principal possible, however, so far only cation exchanged microporous
materials are known.
Several industrial applications of zeolites are based
upon technology adapted from the acid silica/alumina
catalysts originally developed for the catalytic cracking
reaction. This means, that the activity requested is based
on the formation of Brnsted acid sites arising from the
creation of bridging hydroxyl groups within the pore
structure of the zeolites. These bridging hydroxyl
groups are usually formed either by ammonium or
polyvalent cation exchange followed by a calcination
step. The bridging hydroxyl groups, which are pro-

Fig. 2. Formation of Lewis acid sites in zeolites (simplied version


not taking into account the model of true Lewis acid sites) [1].

tons associated with negatively charged framework oxygens linked into alumina tetrahedra, are the Brnsted
acid sites, as demonstrated in Fig. 1 [2].
The protons are quite mobile at higher temperatures,
and at 550 C they are lost as water molecules followed
by the formation of Lewis acid sites, as shown in Fig. 2
[2].
For zeolites, it can be stated that the concentration of
aluminum in the lattice is directly proportional to the
concentration of acid sites. However, for other microporous solids, corresponding correlations are not straightforward [3].
2.2. Carbocations
In general, the nature of acid sites in zeolites is well
understood, however, there is much less consensus on
the reaction mechanisms for hydrocarbon conversion
or formation over microporous materials. It is generally
accepted that the reaction mechanisms of hydrocarbon
conversion and formation on acid zeolites and related
catalysts involve the formation of carbocations. However, whether these carbocations act as transition states
or as intermediates is still under discussion, and is, in
addition, depending on the type of hydrocarbon. The
behaviour of carbocations and their reaction pathways

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

259

Fig. 3. Representation of alkylcarbenium (a) and alkylcarbonium ions (bc). R represents either hydrogen or alkyl group [4]. Reproduced by
permission of Elsevier, Amsterdam.

in zeolites and related microporous materials are


strongly depending on the shape-selective eect due to
the connement of the reacting molecules in the microstructure of the catalystsoering very restricted space
[4].
Concerning the type of carbocations related to the
conversion or formation of hydrocarbons, one has to
distinguish between alkylcarbenium ions (containing a
tri-coordinated positively charged C-atom with three
substituents being either alkyl groups or hydrogens)
and alkylcarbonium ions (consisting of a penta-coordinated positively charged C-atom with the same type of
substituents). See also Fig. 3 [4,5].
The stability of alkylcarbenium ions depends on the
inductive eect of the substituents on the positively
charged C-atom, with the tertiary alkylcarbenium ions
as the most stable ones. However, this eect is less pronounced for the alkylcarbonium ions [4].
In the following, the behaviour of acid sites and the
importance of carbocations in connection with catalytic
conversions using zeolites or related microporous solids
is demonstrated for the cases of aliphatic hydrocarbon
cracking (CC bond scission) and for the alkylation of
isobutane with n-butene (CC bond formation). For
more detailed reviews regarding the reaction mechanisms of acid catalysed hydrocarbon conversions the
interested reader should consult one of the following references [1,4].
2.3. Mechanistic pathways for catalytic cracking of
aliphatic hydrocarbons on zeolites (CC bond scission)
In general, catalytic cracking reactions of hydrocarbons using zeolites can be classied according to the following three main mechanistic pathways:

1. Classical cracking mechanism consisting of a hydride


transfer step to a carbenium ion followed by bscission.
2. Non-classical Haag-Dessau (protolytic) cracking
mechanism proceeding via a carbonium ion transition state.
3. Oligomerisation cracking.
The classical cracking mechanism is based on the fact
that a carbenium ion abstracts a hydride from an alkane
forming another carbenium ion, which cracks by b-scission (cleavage of the CC bond located b to the trivalent
positively charged carbon atom), forming an alkene
see also Fig. 4 [6].
The overall process is governed by the stability of the
carbenium ions in the dierent states of the reaction. In
addition, the reaction rate decreases in the sequence
tertiary > secondary > primary carbenium ions formed.
Furthermore, the activation energy usually increases
with increasing energy level of the nal state. Therefore,
the rate for reactions starting from a tertiary carbenium
ion and ending with a tertiary carbenium ion (type A in
Fig. 5) is faster than the reaction starting from and ending with a secondary carbenium ion (type C in Fig. 5).
The dierent reaction pathways for the b-scission mechanism are summarised in Fig. 5. Please note that these
rather simple assumptions for the b-scission mechanism
correspond quite well with the cracking selectivity observed [1,7,8].
Dehydrogenations are eciently catalysed on the metal sites of bi-functional catalysts, since unsaturated
compounds are much more strongly adsorbed on the
acid sites forming classical carbenium ions than the
saturated ones forming non-classical carbonium ions.
Therefore, classical cracking clearly dominates in

260

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

Fig. 4. Classical cracking mechanism for an alkane molecule [6].


Reproduced by permission of Elsevier, Amsterdam.

bi-functional catalysts. However, classical cracking can


also take place on mono-functional acid catalysts, but
in this case the carbenium ions have to be formed in a
sterically demanding, bi-molecular hydride transfer [9].
Large pore zeolites like Y zeolite show usually a greater
tendency to crack the hydrocarbons according to the
classical cracking mechanism. However, the small and
medium pore zeolites like ZSM-5 favour the non-classical Haag-Dessau mechanism which allow mono-molecular reactions while restricting the bi-molecular (hydride
transfer) reactions due to steric limitations in the pores.
The Haag-Dessau mechanism (see Fig. 6) is the key to
unravel the competing mechanisms of catalytic cracking,
including the classical cracking and oligomerisation

Fig. 6. Non-classical (protolytic) Haag-Dessau cracking mechanism


for an alkane molecule [6]. Reproduced by permission of Elsevier,
Amsterdam.

cracking. Understanding of non-classical cracking has


helped in the diagnosis of shape-selectivity and mass
transfer eects in zeolite-catalysed cracking [6].
From the work of Olah concerning the hydrocarbon
chemistry in superacids it was known that alkanes can
be protonated at low temperatures in liquid phase.
However, Haag and Dessau postulated their mechanism
in 1984 by demonstrating that even zeolites can protonate alkanes to give carbonium ionswhich are transitions states in cracking [10]. The carbonium ions
collapse to give the cracking products: to begin with alkanes (or hydrogen) and smaller carbenium ions, which
then release protons to form the nal cracking product

Fig. 5. b-Scission mechanism for secondary and tertiary alkylcarbenium ions [1,8]. Reproduced by permission of Wiley-VCH, Weinheim.

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

261

Fig. 7. Schematic reaction pathways for carbonium ion decay (protolytic cracking) of a protonated 3-methylpentane together with principle
transition states for dehydrogenation and cracking [1]. Reproduced by permission of Wiley-VCH, Weinheim.

consisting of alkenes [6]. Haag and Dessaus suggestion


is based on the carbonium ion decay of a protonated
3-methylpentane molecule, as shown in Fig. 7.
Since the decay of the carbonium ion leads to breaking of the CC or CH bonds at the insertion point of
the proton, the reaction is called protolytic cracking as
well [1]. Furthermore, compared to the super acid chemistry in liquid phase, the formation of the carbonium
ions using zeolites is only signicant at temperatures
higher than 450 C, and they exist only in a transition
state.
In conclusion, Haag-Dessau cracking (also called
mono-molecular or protolytic cracking) dominates at
low conversions, high reaction temperatures, low reactant pressures and with small and medium pore zeolites
having a low concentration of Brnsted acid sites. All
these conditions favour a low reactant concentration
in the pores and impede hydride transfer. The decay of
the carbonium ion into an alkane and a smaller carbenium ion is the main step in the Haag-Dessau cracking
mechanism [1].

A simplied reaction network for the catalytic cracking of alkanes using zeolites is shown in Fig. 8 [11].
At higher reactant partial pressure the classical cracking mechanism is gradually replaced by oligomerisation
cracking, where we observe substantial oligomerisation
preceding the cracking process. Experimental evidence
for such a route has been demonstrated by Werst et al.
[12] using labeling investigations, and revealing entire
scrambling of carbon-labeled olenic cracking products.
The importance of this mechanism increases with higher
conversion and higher partial pressure as well as lower
reaction temperatures. However, the fundamental chemistry related to this cracking mechanism is basically the
same as observed for the classical cracking mechanism
[1].
2.4. Mechanistic pathway for the alkylation of isobutane
with n-butene on zeolites (CC bond formation)
Zeolites and related microporous materials are also
used as catalysts for the formation of carboncarbon

Fig. 8. Simplied reaction network for the cracking of alkanes on zeolites [11]. Reproduced by permission of Imperial College Press, London.

262

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

bonds, like the very demanding alkylation of isobutane


with n-butene. The reaction pathway of the isobutane/
n-butene alkylation is quite complicated since several
competing reactions can take place besides the regular
alkylation, like self-alkylation, destructive alkylation,
multiple alkylation as well as oligomerisation and cracking. Due to the lower reaction temperature for the aliphate alkylation (thermodynamically favoured), the
desorption step is often dicult in these reactions, since
the reaction product is often more strongly adsorbed
than the reactants. However, the obtained product mixture is an excellent blending component for gasoline,
and the reaction is industrially carried out applying
either HF or sulfuric acid as acid catalysts. There is an
intense search looking for attractive alternatives for
those acids, with large pore zeolites (among other solid

catalysts) as promising candidates, however, so far these


systems suer from an unsuitable catalyst lifetime
[1,4,1318].
The mechanism of the aliphate alkylation can be described as follows: The reaction is initiated by the addition of a proton to the n-butene, forming the secondary
butyl-(2) cation, which abstracts a hydride ion from
isobutane forming tertiary butyl cations. These tertiary
butyl cations interact with n-butene forming isooctyl
cations (preferentially the high octane number representing trimethylpentanes). The isooctyl cations capture
hydride ions from isobutane forming isooctanes and
tertiary butyl cations, which then continue the reaction
cycle (see Fig. 9).
The lifetime of the large pore zeolites (FAU, BEA
and EMT) is determined by the relative rates of hydride

Fig. 9. Mechanism of aliphate alkylation [4]. Reproduced by permission of Elsevier, Amsterdam.

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

transfer and n-butene addition. The latter is usually controlled by using back-mixed reactors operating at
high conversions and low alkene concentrations. The
hydride transfer rate is depending on the stability of
the carbenium ion and the space limitations given by
the microporous framework of the zeolite. The slower
the hydride transfer the more multiple alkylation takes
place, forming C12 and C16 units, which block the active
sites as well as the micropores and deactivate the catalyst [1].
Typical side reactions are classical cracking and oligomerisations over weak Brnsted acid sites, producing
larger olens, which participate in alkylation and lead to
the formation of larger alkylate molecules, which again
contribute to increased deactivation of the catalyst [1].
The coke is formed through a combination of hydride
transfers, inter- and intra-molecular alkylation reactions
and oligomerisations leading to heavy, unsaturated cyclic and acyclic compounds [4].
2.5. Shape selectivity of acid zeolites and related
microporous materials
In 1960, Weisz and Frilette [19] introduced the
expression Shape-Selective Catalysis by demonstrating that Ca A zeolite dehydrated 1-butanol at 260 C
but not isobutanol. This observation showed that the
conversion took place inside the microporous structure
of the Ca A zeolite (0.5 nm pore diameter), not available
for the branched isobutanol due to its large kinetic
diameter. This size exclusion model has since then been
used to remove linear hydrocarbons from mixtures containing both branched and linear hydrocarbons.
The mechanisms of molecular shape-selective catalysis can be described and summarised as follows:
Reactant selectivity describes the phenomenon of
microporous catalysts acting as molecular sieves and
excluding bulky molecules from entering the intra-crystalline void-structure while allowing smaller molecules
to enter. The critical exclusion limit can be varied over
a wide range of dierent zeolites and related microporous solids [4].
Product selectivity refers to discrete diusivities of the
reaction products formed with respect to the microporous pore architecture and crystal size of the catalyst
particles. Sterically less hindered product molecules
may easily leave the microporous framework, whereas
bulky product molecules may stay much longer in the
cavities of the zeolites. The term molecular trac control
has been coined by Derouane and Gabelica in 1980
describing qualitatively the transport of molecules
with dierent shape and/or size in the microporous
framework of zeolites with two discrete sets of pores
[4,20].
Restricted transition state-type selectivity occurs when
the spatial conguration around a transition state or a

263

reaction intermediate located in the intra-crystalline


volume is such that only certain congurations are possible. This means the formation of reaction intermediates and/or transition states is sterically limited due to
the shape and size of the microporous lattice allowing
the access of the species formed to interact with the active sites. This type of selectivity was rst proposed by
Csicsery [21] and is usually connected to the suppression of undesired side reactions like coke formation.
Whereas the product selectivity depends on the crystal
size of the catalyst the restricted transition state-type
selectivity is not depending on the relative rates of diffusion and reaction, hence both selectivities can easily
be distinguished by changing the crystal size of the
catalyst [22].
The dierent types of shape-selectivities of zeolites
and related microporous materials are summarised in
Fig. 10.
Zones et al. [23] introduced the term inverse shapeselectivity for those cases where the restricted transition
state-type selectivity arises from a positive discrimination of specic transition states. An example is the skeletal multiple branching of n-hexane in large pore zeolites
or related microporous solids, where the highest selectivity for multiple branched isohexanes was registered for
microporous solids with well dened pore diameters
and with an optimum interaction with the desirable
isomers [4]. See also Fig. 11.
The cage or window eect, observed in connection
with the hydrocracking of long n-alkanes, represents a
certain case of molecular shape-selectivity in zeolites.
This eect is in operation when the diusivities and/or
reactivities do not change monotonically within a
homologous series of compounds, due to the fact that
certain cracking products, which t the cage dimensions
and are trapped in the cage of the zeolites, were not observed as products [4]. Gorring reported for the rst
time this eect in connection with the hydrocracking
of hexadecane using erionite [24]. Only small amounts
of C7C9 alkanes were observed (although representing
the central cracking products of the probe molecule
and detected in large amounts when cracking hexadecane without using a shape-selective catalyst), since they
t excellently within the cage of erionite [4].
Selective reactions at the pore mouth of zeolites have
been observed by Martens et al., for example the longchain n-alkane isomerisation over Pt/H ZSM-22 (TON
structure) revealing large amounts of mono-branched
isomers although these isomers cannot desorb from
the narrow channels of this mono-dimensional zeolite
[2528]. This observation has been termed pore mouth
catalysis since the product pattern is explained by
involving only the acid sites at the entrance of the small
pore zeolites such as Theta-1, ZSM-22 and erionite. In
addition, the second branching of n-alkanes over Pt/H
ZSM-22 was shown to occur at approximately the

264

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

Fig. 10. Schematic representation of the three types of shape-selectivity. Concerning the restricted transition state-type selectivity: the lower
transition state molecule is easier to accommodate in the cavities than the upper one [1]. Reproduced by permission of Wiley-VCH, Weinheim.

Fig. 11. Inverse shape selectivity observed for the skeletal multiple branching of n-hexane in large pore zeolites. AFI microporous solids represent the
optimum pore diameter [4]. Reproduced by permission of Elsevier, Amsterdam.

distance between the pore openings at the surface of the


zeolite crystallites [22]. This phenomenon has been
termed key-lock catalysis [26].

A schematic presentation of the latter type shapeselective eects in zeolites and related microporous solids is given in Fig. 12.

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

265

Fig. 12. Shape-selective environments in dierent zeolite structure types: (a) large molecules have access to interrupted cavities and channel
intersections for pore mouth catalysis; (b) molecules are plugged into the pore aperture; (c) molecules are converted in multiple pore mouths
according to key-lock catalysis; (d) molecules are converted in the intra-crystalline shape-selective environment [4]. Reproduced by permission of
Elsevier, Amsterdam.

3. Hydrocarbon conversion and formation, crude oil and


natural gas upgrading
3.1. World market of zeolites and related microporous
materials with focus on their application as catalysts
Zeolites and related microporous solids are important components of adsorbents and catalysts used in
connection with the upgrading of crude oil and natural gas, among others due to the fact that they are
easily separated from the educts and products and
by representing a clean technology compared to their
predecessors.
The world market of zeolites and related microporous solids is still in a period of strong development.
Currently about 1.6 millions of tons are used per year,
of which about 1.3 millions of tons refer to synthetic
zeolites and about 0.3 millions of tons to natural zeolites, the latter mainly applied as adsorbent and ion exchanger [29,30].
Concerning the application of synthetic zeolites and
related microporous materials, the focus in terms of
amounts is denitely on detergent builders (1.05 millions
of tons per year), followed by catalysis (0.15 millions of
tons per year) and nally adsorption (0.1 millions of
tons per year). A-type zeolites are by far the most commonly applied detergent builders. Furthermore, A-type
zeolites are mainly used with respect to the application
related to adsorption, separation and purication, which
covers, among others, insulating windows, purication
of olens, natural gas as well as industrial gas, desiccation of alcohols, separation of parans and xylenes
and, nally, production of oxygen and hydrogen. X-type

zeolites are applied as adsorbents for the elimination of


trace amounts of polar impurities, whereas highly siliceous mordenite and ZSM-5 are used for desiccation
of acid gases and the elimination of volatile organic
compounds [29,30].
Finally, almost all the zeolites and related microporous solids, which are used as catalysts, are applied in
the upgrading of oil and natural gas, that means oil rening and petrochemicals. Within oil rening, the main
applications are uidised catalytic cracking (FCC),
hydrocracking, C5/C6 isomerisation and dewaxing,
whereas in petrochemicals, the principal applications
are related to the dierent transformations of aromatics (alkylation, transalkylation, isomerisation, . . .). The
Y-type zeolite present in the FCC catalysts accounts
for almost 95% of the total world consumption of zeolites used within catalysis [30]. Table 1 shows an overview
of the zeolites and related microporous materials used as
catalysts in dierent modied forms on an industrial or
pre-industrial scale in connection with the corresponding
processes [3033].
3.2. Crude oil and natural gas upgradingpresent
scenario and the future
Handling all aspects of crude oil and natural gas
upgrading and their impact on corresponding catalyst
development requires an analysis of the current situation
as well as an evaluation of the major driving forces in
order to meet the future requests from an economic,
technological and environmental point of view.
Modern oil reneries use crude oil of various origins
and they have to meet dierent market demands

266

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

Table 1
Overview of zeolites and related microporous materials used as catalysts in dierent modied forms on an industrial or pre-industrial scale in
connection with the corresponding processes [30,33]
Zeolite/microporous material

Process or application technology

LTA (A-type zeolites)


FAU (X- and Y-type zeolites)
BEA (Beta zeolite)
MOR (Mordenite)

Detergent builder, separation, desiccation


Catalytic cracking, hydrocracking, separation, purication and desiccation, aromat alkylation
FCC additive, cumene and ethylbenzene production
Hydrocracking, hydroisomerisation, dewaxing, NOx reduction, adsorption, cumene synthesis,
transalkylation of aromatics
Ethylbenzene and cumene production
Dewaxing, hydrocracking, ethylbenzene (Mobil-Badger) and styrene production,
xylene isomerisation, methanol to gasoline (MTG), benzene alkylation, adsorption,
catalytic aromatisation, FCC additive, toluene disproportionation
Selectoforming, hydrocracking
Catalytic aromatisation
Methanol to olens (MTO)
n-Butene skeletal isomerisation
Long-chain paran isomerisation
Long-chain paran isomerisation

MWW (MCM-22)
MFI (ZSM-5)

ERI (Erionite)
LTL (KL-type zeolites)
CHA (SAPO-34)
FER (Ferrierite)
TON (Theta-1, ZSM-22)
AEL (SAPO-11)

depending on which country or part of the world they


are serving. Consequently, the process facilities at dierent reneries can vary, however many processes can be
implemented. The market demand is mainly related to
gasoline, diesel, kerosene and fuel oils, which have to
meet specications which are requested by national governments or the European Commission [34]. As an
example, the most advanced category of the car industrys World-Wide Fuel Charter (WWFC) concerning
gasoline specications is summarised in Table 2 [35].
The quality of gasoline is usually dened by the motor octane number (MON, built up with isoparans and
ethers and reduced by the presence of alkenes) and the
research octane number (RON, obtained through the
presence of aromatics and ethers). Diesel is mainly characterised by the cetane number, which is linked to the
presence of alkanes. The cetane number is decreased
by the presence of higher aromatics [34].
World-wide are about 40 million barrels (one barrel
corresponds to 159 l) crude oil rened very day in the
reneries. The basic processes for rening crude oil are
still the same but the tendency is in the direction of more
complex process technology. To begin with crude oil is
divided into various fractions by atmospheric distilla-

Table 2
Most advanced gasoline specications proposed in the World-Wide
Fuel Charter (WWFC) [35]
Parameter
Low octane gasoline [(MON + RON)/2]
High octane gasoline [(MON + RON)/2]
Oxygen (wt.% max.)
Benzene (vol.%)
Aromatics (vol.%)
Olens (vol.%)
Sulfur (ppm)

86.8
93.0
2.7
1.0
35.0
10.0
510

tion. During this procedure the main fractions of oil


products are obtained, which cover
Liqueed petroleum gas (LPG)
Naphtha
Middle distillates
Diesel/gas oil
Lube base oils/atm. residue

C1C4 cut
C5 to about 180 C
130300 C
150370 C
higher than 370 C

The residue of the atmospheric distillation can be


used as feed for a vacuum distillation, leading to the
fractions termed vacuum gas oil (VGO, 370540 C)
and vacuum residue (higher than 540 C).
All these fractions are the primary oil products,
however, many of them have to be upgraded before
they can meet the requested specications and be used
as commercial products [36]. In the case of large
amounts of gasoline to be produced, a FCC unit will
be installed (feed stocks: VGO, residues). Furthermore,
polymerization of the light alkenes obtained or their
alkylation with isobutane is implemented in the renery. Isomerisation of C5/C6 n-parans may be installed,
if there is a need for an increase in the octane number
(getting branched parans). If there is a stronger need
for diesel oil a hydrocracker unit may be added, and if
the residue is too viscous to be handled, a visbreaker
unit may be requested. The reneries apply a lot of catalytic units in order to upgrade, to convert or to purify
their product streams, only visbreaking and coking are
thermal processes. Only a few catalytic units use liquid
catalysts, like the isobutane/n-butene alkylation (HF or
H2SO4), the conversion of mercaptans into disuldes
(Co-phthalocyanines) or alkene dimerisation (Ziegler
Natta type catalysts). All other catalysts are solids,
and the processes applying zeolites or related microporous materials will be presented and discussed in the
following sub-chapters [34]. However, a presentation

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

of a modern renery is not complete without taking


into account the need of the petrochemical industry
which requests synthesis gas, olens and aromatics
(benzene, toluene and xylenesBTX) as basic feed
stocks. Olens and aromatics are usually manufactured
from naphtha via steam cracking [36], however, the
increasing demand of propene can only be satised by
using other technologies like deep catalytic cracking
(DCC) and dehydrogenation of propane in addition
to the steam cracker production. Furthermore, a new
route has been introduced by Weitkamp et al. converting surplus aromatics from pyrolysis gasoline into high
value steam cracker feed (ethane, propane and n-butane) through a hydrodecyclisation step using a bi-functional Pd/H-ZSM-5 catalyst [37]. This technology
combines a considerable decrease of the requested benzene gasoline content with an improved utilisation of
surplus aromatics by increasing the yields and selectivities of ethene and propene in the steam cracking process and represents a real break-through with respect
to high value basic products within renery and
petrochemistry.
The stagnation of crude oil reserves and the increase
of their prices have recently driven the attention towards
the production of fuels and chemicals from natural gas
via synthesis gas (syngas). This route is also known as
the gas to liquids (GTL)-technology. Fuels production from syngas (in former times obtained from coal)
has been reported by Fischer and Tropsch in 1923 for
the rst time [38], using an alkali-promoted iron catalyst. Fuels manufactured via the FischerTropsch route
reveal an excellent quality since they consist mainly of
linear parans and a-olens and do not contain sulfur
and aromatics. A Co-containing catalyst is applied for
the production of heavy parans via the FischerTropsch route starting with natural gas, a technology developed by Shell and named the Shell Middle Distillate
Synthesis (SMDS) route [39,40]. Finally, diesel fuel
(or gasoline) is produced by hydrocracking of the more
or less sulfur and nitrogen-free wax obtained through
the SMDS process using noble metal containing zeolites.
The more restricted fuel specications currently introduced in order to reduce the environmental impact of
hazardous emissions represent a driving force with
respect to an increased use of fuels prepared via the
FischerTropsch route as a blending component of the
gasoline and diesel pools in the future [39]. Besides
the SMDS technology an alternative has been presented
by SASOL/Chevron termed as the Slurry-Phase-Distillate process, again based on the FischerTropsch route
producing wax (using a Co-containing catalyst) followed by a hydrocracking step in order to get diesel or
gasoline [41]. Finally, besides the dehydrogenation of
propane and the DCC process other alternatives have
recently been introduced with respect to meet the
increasing demand of propene, like the Methanol to

267

Propene (MTP) process of Lurgi applying H-ZSM-5


based catalysts [42].
The following sub-chapters will focus on the current
technology and future developments related to dierent
processes dealing with upgrading of crude oil and natural gas and applying zeolites or related microporous
solids as catalysts.
3.3. Isomerisation of n-parans
3.3.1. C5/C6 isomerisation (light straight
runLSRnaphtha)
Environmental restrictions have caused the phase-out
of lead additives and the elimination (or lowering) of
benzene from gasoline, with the consequence of an increased demand of isoparans in order to improve the
octane number of gasoline. Therefore, isomerisation of
light straight run naphtha, containing C5/C6 n-parans,
has advanced to be an important process in the oil
renery. Skeletal isomerisation of n-parans is an
acid-catalysed and equilibrium limited reaction, which
is thermodynamically favoured at lower temperatures,
see Fig. 13 [39].
Industrially, the C5/C6-isomerisation is performed
using a bi-functional catalyst (noble metal together with
an acidic carrier) and in the presence of hydrogen. The
main advantage of applying a bi-functional catalyst is
that stable operations are possible under a suciently
high hydrogen pressure. The mechanism of this reaction
is well accepted, and can be summarised as follows (see
also Fig. 14):
1. The n-parans are initially dehydrogenated on the
noble metal sites to give the corresponding n-alkenes.
2. The n-alkenes are then protonated at the acid sites,
resulting in carbenium ions.

Fig. 13. Thermodynamic equilibrium for hexane isomerisation [36].


Reproduced by permission of Elsevier, Amsterdam.

268

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

Fig. 14. Mechanism of the n-paran isomerisation [36]. Reproduced by permission of Elsevier, Amsterdam.

3. The carbenium ions undergo a skeletal isomerisation


to form more stable branched carbenium ions, following the protonated cyclopropane (PCP) intermediate mechanism suggested by Brouwer [43].
4. Finally, the branched carbenium ions are hydrogenated at the noble metal sites and desorb as isoparafns [39].
Fig. 14 suggests that isomerisation and cracking may
occur as parallel reactions in the presence of hydrogen,
in addition to sequential cracking of pre-isomerised
compounds and post-isomerised cracking products.
Since cracking and isomerisation are both catalysed by
similar acid catalysts, it is not surprising that cracking
takes place besides isomerisation. Skeletal n-paran
isomerisation requires at least a chain of four carbon

atoms, whereas cracking requests a minimum of seven


atoms in the carbon chain. That means that pentanes
and hexanes easily can be isomerised but not easily
cracked. For parans higher than hexanes, cracking is
usually a competing reaction to skeletal isomerisation,
resulting in a lower selectivity for the isomerisation reaction, in spite of the fact that isomerisation is carried out
at lower temperatures than cracking [44]. Weitkamp
demonstrated the dierent results on isomerisation and
cracking of C6C10 n-parans using a bi-functional Pt/
Ca zeolite Y. See Fig. 15 [45].
Commercial isomerisation catalysts contain both a
noble metal (Pt) based hydrogenation-dehydrogenation
function and an acid function. The acid function is provided by either a halogenated (Cl, F) alumina carrier, a
sulfated zirconia substrate or by a zeolite (usually mord-

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

Fig. 15. Comparison of the hydroisomerisation and hydrocracking of


C6C10 n-parans over Pt/Ca zeolite Y [45]. Reproduced by permission of Elsevier, Amsterdam.

enite (MOR)). A zeolite omega (MAZ structure type)


based catalyst has been demonstrated to be superior to
the mordenite based system (higher activity and better
selectivity), however, no commercial use has been reported so far. The higher yield achieved by using a Pt/
H-MAZ catalyst has been reported to be due to both
the unique structural properties of MAZ and the higher
acid strength of its Brnsted sites [46]. Besides mordenite
(MOR) and zeolite omega (MAZ), Pt/Beta zeolites
(BEA) have been investigated as isomerisation catalysts
as well, however, the Pt/H-MOR is the only system
which has been commercialised so far [39].
The halogenated alumina and sulfated zirconia based
isomerisation catalysts are more strongly acidic than the
zeolite based catalysts, which means that they can
isomerise the LSR naphtha at temperatures below
150 C. Consequently, this favours the formation of
the desired isoparans due to the thermodynamic conditions. On the other hand, the tolerance level of the
non-zeolitic catalysts against water and sulfur is not
high, and this leads to a fast deactivation of the halogenated alumina and sulfated zirconia based isomerisation
catalysts and to the request of a severe pre-treatment of
the naphtha feed. Finally, a continuous stream of halogen has to be added in order to keep the catalyst active,
leading to corrosion problems for the reactor system
[39]. The zeolite based isomerisation catalysts are also
lacking sulfur tolerance, however, to a much lesser extent [46].

269

The Pt/H-MOR catalysts are less acidic than the


Pt/halogenalumina catalysts, and, consequently, they
have to be applied at higher reaction temperatures
(about 250 C), which limits the formation of isoparafns due to the thermodynamic conditions. However,
they are more robust than the non-zeolitic catalysts
and can withstand low levels of impurities such as sulfur
and water in the feed [36,39].
The most known example of a zeolitic isomerisation
catalyst is the Pt/H-MOR catalyst, rst developed for
the Shell Hysomer process. The process operates at
2730 bar hydrogen pressure and a reaction temperature
of 250 C. At this temperature, not all normal parans
can be converted to branched parans (see also Fig. 13).
Therefore, it seems to be attractive to combine the
Hysomer isomerisation process with the ISOSIV iso/
normal paran separation process (using a Ca A zeolite
as selective adsorbent for the n-parans), commercialised by Union Carbide (now UOP). The combined technology is known as total isomerisation process (TIP)
and commercialised by UOP (see Fig. 16). The octane
gain in the TIP process is reported to be in the range
of about 10 octane numbers [46]. UOPs zeolitic isomerisation catalyst (Pt loaded mordenite) is commercialised
under the trade name HS-10.
The stability of the Hysomer catalyst is also demonstrated by its long lifetime: catalyst charges have been
used for up to seven years in commercial operation, however, a catalyst deactivated by operational mishaps can in
most cases be regenerated by a simple coke burn-o [44].
An important parameter controlling the isomerisation activity and selectivity is the framework Si/Al ratio
of the zeolite. A maximum of activity for n-pentane
isomerisation was observed for a Si/Al ratio of about
10, for which all framework aluminium atoms are isolated, and therefore, supporting the strongest framework Brnsted acid site possible within the mordenite
structure [4749]. Increasing the lattice Si/Al ratio by
dealumination is benecial as well with respect to catalyst deactivation by reducing the coking rate [50,51].
Dealumination by acid leaching decreases the number
of Brnsted acid sites and creates mesoporosity inside
the zeolite crystallites, which again leads to a decrease
of the diusion limitations. The mesoporosity causes
shorter residence times and facilitates easier desorption
of the products, avoiding secondary reactions of the
intermediates formed and improving the overall selectivity [46]. Besides the nal lattice Si/Al ratio, the method
of dealumination applied has an inuence on the catalyst activity: steam dealumination leaves the lattice
aluminium removed in extra-framework positions
(EFAl), whereas acid leaching results in almost EFAlfree samples. In addition, acid leaching forms aluminium gradients along the zeolite crystallites, whereas
steam treatment produces a more uniform aluminium
distribution. The most active Pt/H-MOR has been

270

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

Fig. 16. Paran Total Isomerisation Process (TIP) [36]. Reproduced by permission of Elsevier, Amsterdam.

obtained by dealuminating the zeolite through acid


leaching followed by mild steaming, leading to small
and controlled amounts of EFAl which create a synergetic eect on the Brnsted acid sites associated to
framework aluminium (FAl), improving their overall
acid strength [39].
CEPSA and Sud-Chemie AG have commercialised
their isomerisation catalyst based on a strongly acidic
Pt/H-mordenite catalyst under the trade name HYSOPAR [46].
3.3.2. Isomerisation of long-chain n-parans
There is a strong interest to extend the isomerisation
reaction to n-parans containing carbon chains longer
than C6, mainly with respect to produce higher octane
multi-branched isomers. However, as mentioned earlier,
one has to take into account that the cracking tendency
of branched parans increases with the length of the
hydrocarbon chain and with the degree of branching.
As an example, Pt/H-MOR produces low yields of isomers in connection with n-heptane isomerisation due
to extensive cracking of the isoheptanes formed [11,39].
Furthermore, isomerisation of long-chain n-parans
has been used to improve the pour point, viscosity,
cloud point and freeze point of middle distillates and
lube oils. In this respect very good catalytic performances have been observed for the n-heptane isomerisation using Pt supported on nano-crystalline Beta zeolite
[52]. This experimental result is explained by a combination of Brnsted acid sites of lower acid strength than in
mordenite and a faster diusion of the branched isomers
through the small crystallites (1020 nm) of the nanocrystalline Beta zeolite, leading to a decrease in the
cracking rate [11].

In addition, besides high isomerisation selectivities


combined with low cracking rates, the degree of branching should be minimized in order to keep a high quality
of the paranic product. In this respect, Pt/SAPO-11 (a
medium pore sized microporous solid) has been demonstrated to display high isomerisation selectivity and low
yields to multi-branched species in the n-octane isomerisation [53]. The suppression of the formation of multibranched isomers on Pt/SAPO-11 catalysts has been
interpreted in terms of a transition state shape-selective
eect induced by the uni-dimensional pore structure of
SAPO-11 [39].
However, in the domain of middle paran isomerisation (C7C9 carbon chain length) there is still a need for
catalyst improvement in order to improve a thorough
isomerisation selectivity (two branches or more) while
minimizing the cracking rate [30].
Isomerisation of higher alkanes in the wax range is
going to play a much more important role in the future
since moderately branched alkanes formed by isomerised/cracked wax represent excellent components for
lube oils. The isomerisation and cracking of synthetic
wax produced through the FischerTropsch route represent an excellent alternative for the manufacture of highquality fuels which are currently derived from petroleum
[44].
3.3.3. Isomerisation of n-butane
Isomerisation of n-butane can take place following
two dierent mechanisms, either via a direct isomerisation involving the formation of a highly unstable
primary carbenium ion (mono-molecular) or via a
dimerisation-cracking mechanism (bi-molecular). The
isomerisation of n-butane is even more thermodynami-

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

cally limited than the C5/C6 n-paran isomerisation.


When running the reaction at low temperatures, very
strong acid sites are requested. In fact, commercial nbutane isomerisation processes make use of HCl/AlCl3
(Phillips [54], Shell [55]) or Pt supported on chlorided
alumina as in the Butamer technology by UOP [56].
Alternative solid acid catalystsmore benecial with respect to the environmenthave been studied, among
others sulfated zirconia, heteropolyacids (in particular
12-tungstophosphoric acid) and zeolites. Sulfated zirconia, however, deactivates very fast. Pt/H-mordenite is
active for n-butane isomerisation, however, this catalyst
requires higher reaction temperatures than sulfated zirconia or heteropolyacids in order to achieve reasonable
conversions. In addition, the zeolite based catalyst forms
larger amounts of dierent by-products than the nonzeolitic catalysts [11,39].
The present n-paran isomerisation capacity world
wide is roughly split equally between n-butane isomerisation, LSR naphtha isomerisation using Pt/H-MOR
and LSR naphtha isomerisation over Pt/Cl/alumina
[44,57].
3.4. Skeletal isomerisation of light n-alkenes
C4 and C5 alkene skeletal isomerisation has been regarded as a suitable alternative for increased production
of isobutene and isopentenes. These isoalkenes are
mainly obtained from FCC units or steam crackers
and are used for the production of methyl tert-butyl

271

and tert-amyl methyl ethers (MTBE and TAME), which


represent excellent fuel oxygenates with good octane
blending properties, in spite of the fact that MTBE
has been questioned as fuel additive for environmental
reasons [58]. In addition, isobutene is also an important
reactant for the petrochemical industry [39].
C4 and C5 alkene skeletal isomerisation is an acid
catalysed reaction requesting strong acid sites. The reaction mechanism can be described by the initial double
bond cistrans isomerisation taking place at the acid
sites before skeletal isomerisation. The protonation of
the double bond leads to the formation of a secondary
carbenium ion, which then rearranges into a protonated
cyclopropane (PCP) structure, and nally ending up
with an isoalkene via the formation of an unstable primary carbenium ion in the case of n-butene, as shown
in Fig. 17 [11].
For thermodynamic reasons, low isomerisation temperatures should be applied, since the equilibrium concentration of branched olens decreases with increasing
temperature. However, at low temperatures the selectivity to isoalkenes decreases due to the competing olen
oligomerisation reactions. The extent of these side reactions can be lowered by applying higher reaction temperatures and low alkene partial pressure, however, at
higher temperatures other non-desired reactions take
place, like cracking, hydrogen transfer and coking, leading to the catalyst deactivation [59].
In former times dierent solid acid catalysts have been
applied as light alkene isomerisation catalysts, like metal

Fig. 17. Reaction mechanism for the acid catalysed skeletal isomerisation of n-butenes [11]. Reproduced by permission of Imperial College Press,
London.

272

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

halides and supported phosphoric acid. Later, environmentally more friendly zeolite catalysts have been introduced successfully, with medium-pore size zeolites as the
most promising systems since the competing (bi-molecular) oligomerisations can be lowered within the limited
space available in the smaller channels of these microporous solids. Among the medium pore sized zeolites those
having a one-dimensional channel structure (like ZSM22 and Theta-1, both having the TON structure, as well
as ZSM-23) turned out to be the best catalysts. They gave
the highest yield of isobutene by working at relatively
low temperatures (below 400 C) and with a low olen
partial pressure. The selectivity to isobutene has been
shown to improve be decreasing the density of the
Brnsted acid sites (increasing Si/Al ratio) and by
decreasing their acid strength by replacing lattice aluminium by other trivalent cations, like gallium and iron. This
behaviour is explained by suppression of bi-molecular
reactions leading to side-products as the acid sites become more and more isolated, and can be regarded as
indirect evidence that isobutene is mainly formed via a
mono-molecular mechanism, as shown in Fig. 17 [11,39].
Finally, a ferrierite based catalyst has been introduced by Shell, giving high yields of isobutene at
350 C and long catalyst lifetimes [60]. This behaviour
has been attributed to the particular structure of ferrierite possessing intersecting 10- and 8-membered ring
channels, which induce the selective formation of trimethyl pentene dimers and their cracking into C4 fragments, including isobutene [61].
In conclusion, quite active and selective alkene isomerisation catalysts can be prepared from microporous
solids by combining the proper pore architecture with
the presence of isolated and/or mild Brnsted acid sites
[39].
3.5. Aliphate alkylation
Aliphate alkylation in connection with oil renery refers to a technology dealing with the partial conversion

of the C4 cut (isobutane and n-butenes) into the socalled alkylate. This alkylate represents a very high quality and valuable component for the renery gasoline
pool due to the high octane numbers of the formed isooctanes, preferentially represented by the trimethylpentanes. Current alkylation technology applies either
hydrouoric (UOP and Phillips) or sulfuric acid (Stracto
and Kellogg). These traditional processes suer from
several safety risks and drawbacks, like the high toxicity,
volatility and corrosiveness of hydrouoric acid and the
high catalyst consumption of sulfuric acid, which requires a regeneration plant close to the alkylation facility. Therefore, replacement of the existing alkylation
processes by new technology based on non-toxic, noncorrosive and environmentally friendly solid acid catalysts is one of the most important research challenges
in the eld of heterogeneous catalysis. Signicant research during the last two decades resulted in a variety
of dierent solid acids capable to produce an alkylate
with the same quality features as the conventionally
manufactured product. However, none of the alternative
solid catalysts have been commercially applied so far,
mainly due to the short catalyst lifetime caused by the
decline of their hydride transfer activity [14,39].
The mechanism of aliphate alkylation has been described in Section 2.4 for the case of isobutane/n-butene
alkylation. Due to typical side reactions like oligomerisation and classical cracking the question arises as to
how long the formed product can be considered as an
alkylate. Arbitrarily, Weitkamp et al. placed this limit
at a content of alkanes in the C8 product fraction of
90 mol.% [62]. In the example shown in Fig. 18, the alkylation stage then ends after about half an hour. This illustrates that the time-on-stream (TOS) behaviour of the
solid catalysts is still unsatisfactorydue to the decline
of the hydride transfer activity. In addition, the ratio between the trimethylpentanes (TMP) and the dimethylhexanes (DMH) formed can be taken as a measure of the
alkylation/oligomerisation ratio for a certain solid acid
catalyst [11]. However, future research within this sub-

Fig. 18. Conversion of a liquid isobutane/1-butene mixture on a CeY zeolite. Composition of the C8 fraction (temperature 80 C, isobutane/1-butene
ratio = 11, pressure: 31 bar) [63]. Reproduced by permission of Wiley-VCH, Weinheim.

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

eld of acid catalysis should concentrate on the reasons


for this loss of hydride transfer activity and measures to
extend the solid catalyst lifetime considerably [14].
In the following, the most recent advances related to
the use of solid acid catalysts for the aliphate alkylation
will be summarised, with focus on the zeolites. An excellent review covering this sub-eld of heterogeneous
catalysis has recently been written by Weitkamp and
Traa, including the early studies with respect to the
search for solid alkylation catalysts [14].
Large-pore zeolites, like FAU, EMT and BEA, have
been in the focus concerning their application as aliphate alkylation catalysts. Extensive studies of the
behaviour of rare earth exchanged faujasites (RE-Y)
were performed by Weitkamp, where he combined a
sampling system downstream of the reactor with highresolution GC in order to obtain detailed results on
the olen conversion and product yields as a function
of time on stream [63,64]. With CeY as catalyst he obtained 100% butene conversion and a high quality alkylate during the rst half hour on stream, after that the
activity decreased rapidly and the selectivity changed
from the alkylate to the oligomerisate (C8 and C12 olens) as the zeolite became deactivated [62].
Corma et al. found a maximum initial conversion of
2-butene for USY zeolites with dierent unit cell sizes
(which means dierent framework compositions) covering a0 values between 2.435 and 2.450 nm [15]. The
TMP/DMH ratio continuously increased with the unit
cell parameter and the hydride transfer activity was
higher for the mildly dealuminated samples [39]. In addition, the extra-framework aluminium (EFAl) formed
during the steam dealumination inuenced the catalytic
performance of USY zeolites in the aliphate alkylation
as well [65].
Cardona et al. investigated the aliphate alkylation
using an USY zeolite at 50 C, concluding with the
observation that their reaction pathway is in line with
the mechanism discussed in Section 2.4 [66]. Gardos
et al. applied rare earth exchanged Y zeolites in the aliphate alkylation using a batch-type autoclave and reaction temperatures between 50 and 100 C, arriving at the
conclusion that the composition of the alkylate is
entirely controlled by the kinetics [67,68].
EMT, the hexagonal faujasite has been investigated
as aliphate alkylation catalyst by Stocker et al., using
a stirred-tank reactor and a reaction temperature of
80 C [6971]. La-H-EMT with a La3+ exchange degree
of 40% was observed to reveal the best alkylation performance, followed by H-EMT and H-FAU and Ce Y zeolite. The better alkylation performance of EMT was
ascribed to a higher strength of the Brnsted acid sites
and slightly larger cages in EMT as compared to the
FAU zeolite [39].
Zeolite Beta (BEA), another large-pore zeolite with a
three-dimensional pore structure, has been used as cata-

273

lyst for the aliphate alkylation [7276]. As for Y-type


zeolites, zeolite Beta deactivated rapidly as well [72].
Corma et al. demonstrated in their more in-depth studies that the performance of Beta as aliphate alkylation
catalyst depends on the synthesis recipe, the crystallite
size, the chemical composition, the post-preparation
treatments, the nature and amount of the extra-framework Al and the density and strength of the Brnsted
acid sites [73,74]. As an example, it was concluded from
these studies that H-Beta zeolite prepared from tetraethylorthosilicate (TEOS) was more active than that synthesised from amorphous silica, and that samples with
crystal sizes of 0.35 lm were more active than those of
0.1 lm [39,73,74]. Finally, Kiricsi et al. [75] and Flego
et al. [76] studied La H-Beta zeolites and the nature of
the carbonaceous deposits formed after adsorption and
conversion of isobutene/1-butene mixtures.
Twelve-membered ring zeolites other than FAU,
EMT and BEA have only found limited attention as aliphate alkylation catalysts, as for example, ZSM-4 (zeolite Omega), ZSM-20 (inter-growth between FAU and
EMT), ZSM-3, ZSM-18 and mordenite [14].
Medium pore zeolites, like ZSM-5 and ZSM-11, have
also been investigated as aliphate alkylation catalysts,
however, these zeolites were found to be active for the
alkylation only at temperatures higher than 100 C,
which is not of interest from a thermodynamic point
of view [77,78]. In addition, these medium pore zeolites
produced less trimethylpentanes, indicating serious pore
restrictions for the formation of the desired alkylate [39].
MCM-22 has been investigated, however, its behaviour
as aliphate alkylation catalyst was found to be in between those of 10 and 12-MR systems [72,79].
Non-zeolitic systems, like sulfated metal oxides, heteropoly acids, supported Lewis and Brnsted acids
and resins have been studied as aliphate alkylation catalysts as well, however, a presentation of these systems
would be outside the scope of this overview. Finally, this
chapter should not be nished before the current process
developments at the pilot plant stage are summarised
concerning the use of alternative solid aliphate alkylation catalysts, however, as far as this is known, no zeolite based systems are among these catalysts [14]:
1. Triuoromethanesulfonic acid on a porous carrier
(Haldor Topse A/S).
2. Antimonypentauoride on acid-washed silica (Chemical Research & Licensing Co., Chevron Corp.).
3. Proprietary catalysts (UOP and Catalytica Inc, Neste
Oy, Conoco Inc.).

3.6. Catalytic reforming


Besides catalytic cracking, catalytic reforming is one
of the most important processes within a modern

274

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

renery, where low-octane naphtha components (normal- and cycloalkanes) are converted into high-octane
isoalkanes and aromatics suitable for gasoline production. The catalytic reforming process consists of a number of hydrogenation and dehydrogenation reactions, in
addition to isomerisation, cyclisation and cracking with
respect to the formation of isoalkanes and aromatics.
The process is run at 425525 C with hydrogen pressures in the range of 0.53.0 MPa [39].
Conventional reforming catalysts are based on Pt/Cl
and PtSn supported on alumina, however, the focus
here will be concentrated on zeolite based reforming catalysts. The non-acidic Linde Type L zeolite (LTL, with
potassium as counter ions) as support for Pt and Ba
has been investigated by several groups [8082]. The
presence of highly dispersed Pt clusters inside the zeolite
channels and the shape-selective eects imposed by the
mono-directional framework structure of the KL-zeolite
is responsible for the very good aromatisation performance of this zeolite, which has a pore diameter of
0.71 nm. One of the main drawbacks of this zeolite is
the high sensitivity towards sulfur poisoning, responsible for the fast catalyst deactivation. The preparation
method and especially the Pt incorporation have a
strong inuence of the catalyst performance with respect
to activity and stability [39].
Other zeolite based reforming catalysts have been
investigated as well, like Pt/ZSM-12, Pt/Beta and sulded Pt/Cs-Beta [83,84]. Finally, large-pore boro-silicate
based zeolites (B-Beta, B-SSZ-33, B-SSZ-24 and B-SSZ31) have been patented by Chevron as reforming catalyst [85].

signicant change in this business occurred in 1942 with


the introduction of the FCC technology. A schematic
diagram of a typical FCC unit is shown in Fig. 19 [39].
The FCC process can briey be summarised as follows: The pre-heated feedstock is contacted with the
hot catalyst coming from the regenerator at the bottom
of the riser reactor, where most of the cracking reactions
take place at temperatures around 500 C and contact
times with the catalyst of about two to three seconds.
The cracking products are hydrocarbons, which are extracted from the catalyst pores in the stripper unit using
steam, and then passed to the regenerator to restore the
catalyst activity by burning o the coke formed during
the cracking reactions at temperatures of about 700 C
[39]. Part of the used catalyst is continuously replaced
by fresh catalyst, which results in a consumption of
about 10 tons catalyst per day for a medium sized
FCC unit. The term uidised refers to the catalyst
particles in the range of 6090 lm consisting of porous
micro-, meso- and macrospheres, which are uidised,
e.g., intimately admixed in a stream of vaporised hydrocarbon feedstock and steam. Since the cracking reactions are primarily endothermic, heat balance with the
exothermic regeneration reaction is required for the riser
to operate at appropriate cracking temperatures. Thus,
the continued operation of an FCC unit depends on
the heat balance of the riser reactor and regenerator [86].
With the introduction of zeolite (faujasite type) containing cracking catalysts in 1962, replacing the amorphous silicaalumina, a tremendous change concerning

3.7. Catalytic cracking


The catalytic cracking unit is the most important conversion facility in a modern renery. This process consists of the scission of the hydrocarbon CC bonds
present in the feedstock (usually vacuum gas oils or residues) in order to obtain gasoline, light alkenes or other
low molecular hydrocarbons. A number of dierent
FCC catalysts exist and catalyst changes in the worldwide about 350 renery FCC units are made often,
depending on the feedstock type and quality available
[86]. This process, which produces about 30% of the
total gasoline pool either directly or indirectly, is very
exible with respect to dierent combinations of process
design and catalysts. This exibility allows the reners
to process a large variety of feedstocks and to adapt
the product pattern to the changing market demands
with respect to local fuel specications and environmental legislation [39].
The history of catalytic cracking started in the 1920
when Eugene Houdry (the father of catalytic cracking)
used an acid treated natural clay as catalyst to convert
hydrocarbons into lower molecular weight products. A

Fig. 19. Schematic diagram of a typical FCC unit [39]. Reproduced by


permission of Wiley-VCH, Weinheim.

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

FCC technology took place. Zeolite containing catalysts


are much more active, show higher gasoline yield performances and produce less coke than the amorphous
silicaalumina based catalysts, resulting in higher feedstock conversions and severities as well as enhanced economic benets of the process [39,86]. The actual FCC
technology involves the formulation of proprietary multifunctional cracking catalysts, consisting of dierent
amorphous (catalytically active macroporous matrix,
clay type binder) and crystalline acid functions (represented by shape-selective (microporous) zeolites like
Y-type zeolite containing mesopores due to dealumination forming the ultra-stable Y zeoliteUSY), and a
series of additives for metal passivation (mainly V and
Ni), sulfur removal, promoters for total combustion
and octane enhancing additives [11]. The two main components of cracking catalysts are the zeolite Y and the
matrix. The matrix plays a critical role in the selective
cracking of the (high molecular) bottoms fractions when
residue containing feedstocks are processed. The main
functions of the matrix are to pre-crack large molecules
and adsorb Ni and V preferentially in order to protect
the zeolite Y of the catalyst particle. In an ideal situation, the pre-cracked large molecules from the matrix
macropores are further cracked in the mesopores of
the USY (to, i.e., gas oil fractions) before, nally, gasoline is formed in the micropores of Y-type zeolite (or
propene in the case of ZSM-5), see also Fig. 20 [87,88].
Concerning metal passivation, both vanadium and
nickel deposit on the cracking catalyst as their host
molecules are converted to lighter products and coke.
Both are extremely deleterious when present in excess
of 3000 ppm on the FCC catalyst. Vanadium in the oxidation state 5+ is converted to vanadic acid and reacts
with the zeolite framework by hydrolysing the zeolite
lattice structure, and, thus deactivating the zeolite part

275

Fig. 20. Conceptual pore architecture design of a FCC catalyst [88].


Reproduced by permission of Elsevier, Amsterdam.

of the catalyst, whereas nickel forms a metal/metal oxide


site on the catalyst surface causing formation of coke
and hydrogen [86,87].
FCC catalysts are nowadays mainly produced by
four companies: Grace Davison, Akzo Chemicals,
CCIC and Engelhard Corporation. The high activity
and relatively low coke formation of zeolite containing
FCC catalysts enabled the reactor technology to advance from dense uidised beds to short-contact-time
(SCT) risers with a corresponding improvement in the
performance [36], see also Fig. 21.
Concerning the mechanism of catalytic cracking, the
acid sites of the zeolite component are regarded as the
catalytically active sites, and the mechanism has been
discussed already in Section 2.3. Most studies in uidised catalytic cracking have focused on zeolite Y, as
this is still the dominant zeolite used in FCC. Besides
the acid properties of this zeolite, the unique pore architecture of Y zeolite is ideal for cracking gas oil components into gasoline molecules. Moreover, it has been
observed that the activity of the Y zeolite for gas oil
cracking has a maximum for a Si/Al ratio of 58, which

Fig. 21. Trends in catalytic cracking catalyst performance [36] Reproduced by permission of Elsevier, Amsterdam.

276

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

corresponds to a unit cell size (UCS), a0, of 2.436


2.440 nm. This clearly indicates that gas oil cracking
requires the presence of strong Brnsted acid sites.
Unfortunately, so far it has not been possible to prepare
Y zeolite with a framework Si/Al ratio above 4 by direct
synthesis. Therefore, highly dealuminated zeolites (low
a0 values) have to be prepared by dealumination of commercially prepared Y zeolite samples with Si/Al ratios in
the range of 2.6. However, it was found that the catalytic
behaviour of dealuminated Y zeolites with given a0 values strongly depend on the method of dealumination applied [89]. In this respect, Y zeolites dealuminated by
steaming (USY) create a secondary porosity formed
during the partial destruction of the zeolite framework
and forming mesopores which facilitate diusion of larger molecules into the zeolitic channels. The obtained
USY type zeolites show, in addition, a much better
hydrothermal stability, which is a pre-requisite of the
application as FCC catalyst (cf. regeneration conditions
of the FCC catalysts) [39,86].
In certain cases the oil companies would like to increase the amount of lighter components, like propene,
n-butenes and isobutene, as they are important feedstocks for the petrochemical industry. ZSM-5 turned
out to be excellent in this respect, especially for the
enhancement of propene. This zeolite can signicantly
improve the octane number of gasoline in catalytic
cracking. Addition of a few per cent ZSM-5 to a conventional FCC catalyst gives an equivalent octane number
increase [90]. Due to the pore architecture, ZSM-5 increases the octane number of the gasoline by selectively
upgrading low octane gasoline components into lower
molecular weight compounds with a higher octane number [11].
The concept of using ZSM-5 as co-catalyst to modify
the performance of a generic FCC catalyst system can
signicantly increase the product exibility in the FCC
unit. An extension of this technology is the so-called
Deep Catalytic Cracking process (DCC), which involves
the use of ZSM-5 as the primary catalyst rather than the
co-catalyst in a FCC type of moving bed reactor system
in order to maximise lower olens production (primarily
propene and a-olens) with gasoline as a by-product
[36,91]. SINOPEC has commercialised this technology
in China and a DCC plant is under construction in Thailand [92]. A modied DCC process has been oered by
SINOPEC, termed as Catalytic Pyrolysis Process (CPP),
in which vacuum gas oils and atmospheric residues are
converted to ethene and propene [93].
Other zeolites, like Beta and MCM-22, have been
investigated with respect to catalytic cracking. They
have shown some use in modifying FCC reactions, however, none of them have managed to balance activity
and product selectivity as well as Y zeolite [94].
Since only 60% of the worldwide demand of propene
can be produced using steam cracking technology, the

remaining amount must be manufactured either by


FCC/DCC or propane dehydrogenation. One future
challenge within catalytic cracking will focus on the enhanced production of light olens by development of the
current technology however, with the continued growth
of residue processing in FCC units, further attention
must be paid to the performance of catalysts for this
variant of FCC operation as well. Since both the zeolite
and the matrix play an important role in the optimal
performance of a residue catalytic cracking catalyst the
following properties have to be considered in detail with
respect to catalyst development [36]:
1. structure of the mesopores for the bottoms
conversion,
2. unit cell size (UCS) for the coke make,
3. resistance of the zeolite structure to vanadium attack
and
4. defect structure with respect to the hydrothermal
stability.

3.8. Catalytic hydrocracking


Catalytic hydrocracking is an oil renery process
which has been developed with respect to the conversion
of relatively heavy oil feedstocks (including residues)
into lighter transportation fuel products through CC
bond scission. In contrast to catalytic cracking, hydrocracking is performed in the presence of hydrogen as
co-feed at relatively high pressures (50200 bar) and
at lower temperatures than catalytic cracking (300
450 C). However, catalytic hydrocracking accounts
for a more hydrogenated product than catalytic cracking, due to the hydrogenation reactions taking place
under these conditions. In addition, the coke formation
rate and the gas yield are considerably lower in catalytic
hydrocracking compared to catalytic cracking [39,95].
Catalytic hydrocracking is a very exible technology,
allowing a wide range of feedstocks to be processed.
Whereas in the US the hydrocracking units are mainly
used to convert lighter feedstocks (straight run light
and heavy gas oils, coker gas oils, FCC cycle oils and
thermally cracked gas oils) into gasoline components,
the hydrocrackers outside the US produce a much wider
spectrum of products, like kerosene, jet fuel and diesel
fuel (also called middle distillates) from vacuum gas oils
(VGO) processing (see also Table 3) [39,95].
Currently, the annual catalytic hydrocracking capacity world-wide amounts to about 200 million tons, distributed over around 120 catalytic hydrocracking units,
which mainly have been developed by UOP/Unocal
(Unicracking process), Chevron (Isocracking and Isomax technology), Shell and IFP [95,96]. Catalytic hydrocracking processes can be performed according to three
main technologies: The single stage conguration, where

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292


Table 3
Hydrocracking feedstocks and products [39]
Feedstock

Product

Straight run gas oils (SRGO)


Vacuum gas oils (VGO)
FCC cycle oils
Coker gas oils
Thermally cracked gas oils
Deasphalted oils
Straight run and cracked naphthas

LPG
Gasoline
Catalytic reforming feeds
Jet fuels
Diesel fuels
Heating oils
Olen plant feedstocks
Lube oils
FCC feedstocks

the feed is processed in a single catalyst bed in one or


two reactors in series, recycling the un-converted feed.
Another conguration employs two catalysts either in
the same reactor (stacked beds) or in two reactors
in series. The rst catalyst is the hydrotreating catalyst
(removal of sulfur and nitrogen), which partially hydrogenates aromatics as well. Catalytic hydrocracking is
then performed on the second catalyst. The two stage
conguration consists of two reactors, where the rst
one contains the hydrotreating catalyst and the second
one the hydrocracking catalyst [39,97]. Finally, a more
modern and cost-eective process is the series-ow conguration, with no product separation in between the
hydrotreating and hydrocracking steps, and, thus, re-

277

quires very robust second stage catalysts such as those


based on zeolites [36].
In general, with increasing feedstock heaviness, the
amount of catalyst poisons (metals, aromatic coke precursors, sulfur and nitrogen) increases. Metals cause
irreversible deactivation of the rst stage hydrotreating
catalyst, while organic nitrogen compounds specically
reduce the cracking activity of the acidic second stage
catalysts [95].
The main reactions occurring during catalytic hydrocracking are summarised in Fig. 22.
Catalysts used in the rst stage for feedstock pretreatment are usually hydrotreating catalysts (Co/
Mo, Ni/Mo or Ni/W supported on alumina or silica
alumina). The real hydrocracking catalysts are bifunctional systems, consisting of a hydrogenation/
dehydrogenation and an acidic cracking function. The
activity and selectivity of the hydrocracking catalyst depends on the ratio between the hydrogenation/dehydrogenation and acid functions as well as on the strength of
both. For example, Lemberton et al. reported the existence of an optimum hydrogenation/acid ratio in Ni/
Mo/zeolite Y/alumina catalysts, resulting in a reduced
amount of coke formation with increasing intimacy of
mixing of the two functions at the submicron level
[98,99]. Amorphous silicaalumina (ASA) is still being
used as acidic carrier in some hydrocracking units [97].

Fig. 22. Basic catalytic hydrocracking reactions [39]. Reproduced by permission of Wiley-VCH, Weinheim.

278

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

However, the incorporation of zeolites into the catalyst formulation during the 1960s represents a real
break-through with respect to an enhanced catalytic performance of hydrocracking catalysts [100]. The zeolite
based hydrocracking catalysts show a higher activity
(due to a higher acid strength), higher thermal and
hydrothermal stability, better resistance to sulfur and
nitrogen containing catalyst poisons as well as lower
coking rate, thus prolonging the catalyst lifetime
[101,102].
The rst generation of Y zeolites (either in their
hydrogen form or exchanged with rare-earth cations)
used in catalytic hydrocracking contained a high
amount of lattice aluminium, which resulted in a high
unit cell size. However, the use of ultrastable Y zeolites
(USY) synthesised by steam dealumination made it possible to control not only the number and strength of acid
sites but also the amount of extra-framework aluminium
as well as the degree of mesoporosity or secondary
porosity. As the severity of the steaming increases, the
density of acid sites decreases along with the unit cell
size. Hydrocracking catalysts synthesised from USY
zeolites with low unit cell sizes (a0 < 2.445 nm) produce
less gas but higher liquid yields and are more selective
towards middle distillates. Furthermore, acid-leached
dealuminated Y zeolites are more active and selective towards middle distillates than parent USY. The presence
of secondary porosity in USY zeolites (due to steam
dealumination) as well as the macroporosity of the for-

mulated hydrocracking catalyst play an important role


with respect to processing heavy fractions and residues.
The mesoporosity or secondary porosity has been
shown to reduce the mass transfer limitations during
hydrocracking and, thus, suppress secondary cracking
(see also Fig. 23) [39,96,103].
Another solution with respect to improvement of the
reactant accessibility and decrease of secondary cracking
reactions has relied on the preparation of small Y zeolite
crystallites (<0.5 lm). Furthermore, the preparation of
mesoporous Al-MCM-41 materials and the application
of their Ni/Mo derivatives with respect to hydrocracking
of vacuum gas oil revealed a good selectivity for middle
distillates and demonstrated a higher activity compared
with Ni/Mo supported on amorphous silicaalumina
[104,105].
Besides Y zeolite, other large pore zeolites or related
microporous and mesoporous materials, such as Omega,
L zeolite, Beta, VPI-5, mordenite, UTD-1, MCM-48
and SBA-15, have been investigated as components for
hydrocracking catalysts, however, these systems have
not yet led to commercial application [39,96,106].
Recent developments within hydrocracking catalysis
have been concentrated on the improvements in the
amorphous silicaalumina and Y zeolite base materials
and the control of catalyst pore architecture. Although
zeolite based hydrocracking catalysts oer the best prospects for heavy feeds conversion, their activity is suppressed due to the restricted access of the heavier

Fig. 23. Creation of a secondary pore structure in Y zeolite as a result of dealumination [103]. Reproduced by permission of Imperial College Press,
London.

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

molecules to the zeolite pores. Therefore, the main


focus concerning further research will be concentrated
on nding a zeolitic catalyst component dedicated to
the production of middle distillates, associated with the
activity, stability and proper pore architecture of the
nal hydrocracking catalyst.
3.9. Catalytic dewaxing
Heavier oils, like gas oils, diesel oils or lube oils, contain often wax (long-chain normal and/or slightly
branched C18 and higher parans), which crystallise at
low temperatures (below 100 C). This aects their viscosity deleteriously, as quantied by pour point determination [107]. These compounds are therefore often
removed, either by physical processes such as extraction
in solvent dewaxing, or nowadays by catalytic dewaxing
using shape selective catalysts. During catalytic dewaxing (carried out at about 400 C), normal and slightly
branched long-chain parans are removed by selective
cracking to lighter products including gas. The basic
process resembles selectoforming, where C5C9 normal
parans are cracked to LPG. The selective removal of
the long-chain normal parans in gas oil dewaxing
using ZSM-5 as catalyst is illustrated in Fig. 24, where
the sharp peaks of the normal parans in the gas chromatograph of the feed are absent in the product after
catalytic dewaxing [108].

279

Catalytic dewaxing has been connected to the medium pore zeolites and related microporous solids, including ZSM-5, ZSM-11, ZSM-23 and SAPO-11 [109]. BP
introduced the rst generation dewaxing catalysts,
which was based on mordenite in order to remove normal alkanes from lube oils. However, Mobils discovery
of ZSM-5 forced the development of the next generation
of dewaxing catalysts, applicable to both gas oils
(MDDW process) and lube oils (MLDW process). The
new generation of ExxonMobil catalysts (MSDW: Mobil Selective DeWaxing) selectively converts linear longchain parans to their corresponding isoparans
instead of cracking to lower molecules. Studies in this
direction involved zeolite Beta and MCM-22 as catalyst
components. Chevron commercialised their Isodewaxing process, which is based on isomerisation rather than
cracking as well, using SAPO-11 as dewaxing catalyst
[110]. Finally, AKZO-FINA, Criterion/Zeolyst/Lyondell and others oer catalytic dewaxing technology and
catalysts as well [36,95,109].
In conclusion, the most recent developments within
catalytic dewaxing have been concentrated on dewaxing
by means of shape selective isomerisation rather than
cracking. Examples of catalysts which exhibit shape
selective isomerisation properties include SAPO-11
and zeolite Beta. The SAPO-11 catalyst has the advantages of both low cracking activity and good isomerisation activity while suppressing the formation of

Fig. 24. Catalytic dewaxing of gas oil using shape selective ZSM-5 as catalyst [108] (top: gas chromatograph before catalytic dewaxing, bottom: gas
chromatograph after catalytic dewaxing, RT means retention time). Reproduced by permission of Elsevier, Amsterdam.

280

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

multi-branched isomers which, if formed, would easily


crack to undesired lighter products [95].
3.10. Transformation of aromatics/petrochemicals derived
from aromatics
3.10.1. Alkylation of aromatics
Aromatics alkylation represents world wide a large
scale industrial process for the production of intermediates, ne chemicals and petrochemicals. In principle, the
chemical reaction consists in the replacement of a hydrogen atom of an aromatic compound by an alkyl group.
An acid catalyst is required when the replaced hydrogen
is on the aromatic ring (electrophilic substitution), however, base catalysts or radical conditions are required if
the hydrogen on the side chain of an aromatic compound is substituted. Important products manufactured
by aromatics alkylation include ethylbenzene (further
converted to styrene for polymer production), cumene
(isopropylbenzene, an intermediate for the production
of phenol and acetone), alkylnaphthalenes (precursors
to advanced polymers) and alkylbenzene sulfonates
(detergent builders). Acid catalysts applied for aromatics alkylation are usually Brnsted or Friedel-Crafts
acids and cover mineral acids, metal halides, cation exchange resins, acidic oxides and zeolites. Especially the
last group of catalysts is well suited for the specic production of dedicated alkylated aromatic compounds,
including single isomers of those compounds, due to
their discrete pore architecture and Brnsted acidity
[111,112].
Ethylbenzene (EB) is mainly used as a precursor for
styrene monomer. About 90% of the world wide ethylbenzene production is based on alkylation of benzene
with ethylene, with AlCl3 as the primary Friedel-Crafts
catalyst. Due to the corrosive nature of this catalyst,
alternatively, solid acid catalysts have been developed,
like ZSM-5 for the Mobil-Badger vapour phase formation of ethylbenzene, commercialised in the 1970s and
operating at 380450 C and 2030 bar pressure. High
yields of ethylbenzene (more than 99%) can be achieved
and the catalyst deactivation is slow (bi-molecular hydride transfer is largely suppressed due to steric hindrance), leading to long lifetimes applying suitable
regeneration procedures. About 35 plants are operating
world-wide using this technology, with an annual
production capacity of nearly eight million tons. Alternatively, liquid phase alkylation of benzene with
ethylene, using zeolitic catalyst systems have been commercialised, like ExxonMobils EBMAX technology
applying MCM-22 as catalyst [111,112].
Cumene (isopropylbenzene) is manufactured by
alkylation of benzene with propylene, mainly still using
the solid phosphoric acid (SPA) technology (UOP). Zeolites have been investigated extensively as alternative
solid catalysts for the cumene production, including

ZSM-5, mordenite, Beta and ZSM-12. However, so far


only a few technologies have been commercialised using
highly dealuminated mordenite (Dow Chemicals), Beta
(EniChem) or a proprietary catalyst system (ExxonMobil) [111,112].
Gas phase ethylation of toluene with boric or phosphoric acid modied ZSM-5 yields para-ethyltoluene
with up to 100% isomeric purity, which is an important
precursor for manufacturing para-methylstyrene (advanced polymer production). ExxonMobil and Deltech
Corp. have commercialised this technology [113].
Cymene (isopropyltoluene) is manufactured commercially by alkylation of toluene with propene using solid
phosphoric acid catalysts (Sumitomo). Cymene is an
important intermediate in the production of meta-cresol,
and zeolitic materials have been tried as catalysts as well.
Flockhart et al. have applied Y zeolite to perform this
reaction [114].
Side-chain alkylation of toluene with methanol has
been performed using Cs-exchanged X zeolite as base
catalyst in order to produce ethylbenzene and, subsequently, styrene [115]. This technology is close to be
commercialised [93].
Alkylation of naphthalene with propene over zeolite
catalysts yields mainly 2,6-diisopropylnaphthalene
[116], whereas the alkylation of naphthalene with methanol can either yield preferentially 1-methylnaphthalene
(ZSM-12, 12-MR system, kinetic control) or 2-methylnaphthalene (dealuminated mordenite or ZSM-5) [111].
3.10.2. Isomerisation and transalkylation
of alkylaromatics
The transfer of alkyl groups between aromatic molecules, also termed transalkylation, and the intra-molecular isomerisation are commercially applied in large
scale. Both reactions are acid catalysed processes, and
most of the catalysts used are solid systems, with zeolites
as the most prominent representatives [117].
The pyrolysis gasoline from naphtha crackers and the
naphtha reformate consist mainly of xylenes and ethylbenzene with respect to their C8 aromatics fractions.
They are isolated from these streams by distillation
and solvent extraction. Xylenes are used on a large scale
industrially, and they are precursors for a number of
important petrochemicals. Especially para-xylene has
an enormous market potential (terephthalic acid production for polyester formation), with an annual increase of about 7% [93,118].
When C8 aromatics fractions are processed with respect to xylene isomerisation, the remaining ethylbenzene must be converted to xylenes. Catalysts used for
the xylene and ethylbenzene isomerisation contain always platinum, in former times mainly on chlorinated
aluminas or steamed silicaaluminas (Octaning process
developed by Atlantic Rening Corp.), nowadays preferentially on mordenite. The process is conducted in a

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

xed-bed reactor at temperatures between 370 and


430 C and a pressure range of 715 bar. New generations of zeolite based catalysts have been introduced
by UOP (Isomar process) and IFP (Oparis), claiming
even higher yields of para-xylene. Metal modied
H-ZSM-5 catalysts are most eective when the xylene
isomerisation is accompanied by ethylbenzene dealkylation to benzene and ethene under similar process conditions (ExxonMobil) [118].
Disproportionation is a special case of transalkylation and is in place when alkyl groups are transferred
between identical molecules. Several processes are
commercialised dealing with toluene disproportionation
using mordenite or ZSM-5 as catalysts and arriving at
thermodynamic mixtures of xylenes (ExxonMobil,
UOP and TotalFinaElf). Selective toluene disproportionation processes are on the market as well, using
ZSM-5 catalysts and arriving at para-xylene rich product mixtures with selectivities of more than 80% (EniChem, ExxonMobil and UOP) [117,118].
The disproportionation of ethylbenzene to benzene
and diethylbenzenes has been studied extensively
[119,120], but has not been commercialised until recently. With large pore zeolites (Y zeolite), the reaction
occurs via a hydride transfer chain reaction through
diphenylethanes as intermediates. Medium pore zeolites
(ZSM-5) cannot accommodate this bulky intermediate,
and ethylbenzene disproportionation proceeds via a
dealkylationrealkylation path. This technology is commercialised using a modied ZSM-5 catalyst [117]. The
Catalysis Commission of the International Zeolite
Association (IZA) has recommended the disproportionation of ethylbenzene using LaNaY zeolite as a standard reaction for acidity characterisation of acid
zeolites [121].
3.10.3. Conversion of surplus aromatics
In connection with the introduction of the European
Auto Oil Programme, the aromatics content of gasoline
has to be reduced currently from 43 to 35 vol.% in 2005.
This number is also expressed by the most advanced gasoline specications proposed in the World-Wide Fuel
Charter (WWFC, see Table 2) [35]. As a consequence,
there will be a surplus of aromatics. The main source
of aromatics is the so-called pyrolysis gasoline, a byproduct rich in aromatics from the manufacture of ethene and propene by steam-cracking of straight run
naphtha (light hydrocarbons). Due to the predicted
world-wide increasing demand of ethene and propene,
the surplus of pyrolysis gasoline will increase further
[122]. A novel catalytic process for hydrogenative ring
opening of aromatics has been introduced, which allows
the conversion of pyrolysis gasoline from naphtha
steam-crackers into a high quality synthetic steam
cracker feed composed of C2C4 n-alkanes [37,123].
There are two process variants, the direct conversion

281

of aromatics with hydrogen on bi-functional zeolites


and the two-stage process comprising a conventional
ring hydrogenation to cycloalkanes followed by ring
opening of these compounds using acidic zeolites. A
large part of this research has been carried out in the
laboratories of J. Weitkamp at the University of Stuttgart (Germany) [37,123127].
The main advantage of the direct route is the fact that
one single reactor is sucient to perform the desired catalytic conversion, which is carried out using shape selective bi-functional zeolites, like Pd/H-ZSM-5 (Si/Al ratio
of about 20). For example, ring hydrogenation (6 MPa
hydrogen pressure) of toluene to methylcyclohexane
and skeletal isomerisation of the cycloalkane into ethylcyclopentane and dimethylcyclopentanes occurs at temperatures up to about 250 C, however, ring opening
with respect to the desired formation of C2 n-alkanes
(73% of the total synthetic steam cracker feed) requires
temperature of up to 400 C [124]. C2 n-alkanes refers
in this context to ethane, propane and n-butane.
Advantages of the two stage route are an easier removal of the heat generated in the two exothermic steps
and the possibility to optimise the reaction conditions of
ring hydrogenation (performed by conventional technology using metal catalysts) and ring opening (carried out
by novel shape selective zeolites in their acid form as catalysts). Applying H-ZSM-5 with a Si/Al ratio of 20,
yields of C2 n-alkanes, comparable to those obtained
during the direct conversion of toluene on Pd/H-ZSM5, are achieved when processing methylcyclohexane at
400 C as second step of the two stage route. ZSM-5
and ZSM-11 turned out to be the best suited catalysts
for this technology, whereas large pore zeolites (like
Beta or Y zeolite) deactivated rapidly. Zeolites with
too narrow pores, like ZSM-35 or ZSM-22, did not
show high degrees of ring opening capacities, probably
due to the hindered diusion of the hydrocarbons involved [124].
3.11. Aromatisation of light alkanes
Aromatics (BTX) are precursor compounds for the
petrochemical industry of large importance. Since the
liquied petroleum gas (LPG) fractions in the renery
are supposed to increase due to more severe operations
of the FCC units, there is a need to convert the low
value LPG into aromatics.
The aromatisation of light alkanes is easier to perform as the size of the alkane increases, taking into
the thermodynamic relations. For example, propane
and higher alkanes can be aromatised at temperatures
lower than 500 C, whereas temperatures of up to
575 C are requested for the aromatisation of ethane.
In addition, the proportion of benzene in the BTX fraction tends to decrease with increasing size of the alkane
[11,39,128].

282

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

The mechanism of the light alkane aromatisation involves the initial dehydrogenation of parans to form
lower olens and hydrogen, followed by olen oligomerisation and cyclisation reactions to produce aromatics,
in which hydrogen transfer plays an important role
[36], see also Fig. 25.
Several zeolite based industrial processes have been
introduced with respect to the aromatisation of light
alkanes. The most known process is BP/UOPs CYCLAR process using C3/C4 renery gases as feedstock
and Ga/H-ZSM-5 as catalyst (15 wt.% Ga, Si/Al ratio
between 15 and 30). However, other technologies are
known, including Mobils M2 Forming process (using
H-ZSM-5), Chevrons AROMAX process applying
C6C8 alkanes as feed and Pt/L zeolite as catalyst as well
as Selectoforming over erionite [128]. An advantage for
the CYCLAR process is the fact that signicant
amounts of hydrogen are formed, which is needed in
the renery and which makes this process economically
very attractive [11].
The preferred catalyst for alkane aromatisation is Ga/
H-ZSM-5. The incorporation of Ga can be performed
either by direct synthesis, impregnation or ion exchange.
The success of the modied H-ZSM-5 catalyst is connected to the limited coke formation by using this
zeolite. Ga modication improves this behaviour even
more [128]. Besides gallium, other metals such as Zn,
Pt, Ni and Ag have been used in combination with
ZSM-5 for the light alkane aromatisation, however,
the Ga- and Zn/H-ZSM-5 modied catalysts are the preferred systems. Finally, Pt/KL zeolite has been applied
in the selective formation of benzene from hexane.
There is a strong interest in the aromatisation of ethane since this compound is an important component of

both renery and natural gases. However, from a thermodynamic point of view, it is very dicult to convert
ethane into aromatic hydrocarbons, since a highly
unstable primary carbenium ion will be formed in the
acid catalysed oligomerisation step, whereas for higher
alkanes a more stable secondary carbenium ion is
formed. Anyway, several investigations have been reported using Pt, Pd, Ga or Zn modied H-ZSM-5 zeolites and temperatures of about 575 C [39,128130].
3.12. Natural gas upgrading/gas conversion technologies
The evolution of the known crude oil and natural gas
reserves world-wide indicates a dramatic increase in the
latter compared to a levelling o concerning the crude
oil. This trend is expected to continue, which willin
addition to the price development with respect to the
crude oil based upgradingmost likely generate a gradual shift towards the application of natural gas as a feedstock for the production of fuels and petrochemicals.
This situation has forced an enhanced global interest
in processes, which can convert natural gas into liquids
and higher added value productswithout going via
methanol as intermediate. This route is known as the
gas to liquids (GTL)-technology, based on the
FischerTropsch route. The interest to manufacture
fuels and petrochemicals from natural gas is driven by
the desire to apply this technology directly, for example
at remote natural gas eld sites, in order to minimize
transportation costs and gas burning at the recovery
sites [36].
The following sub-chapters will deal with the
GTL-technology, based on the FischerTropsch
route as well as the methanol to hydrocarbon conver-

Fig. 25. Simplied reaction mechanism for the aromatisation of propane over modied H-ZSM-5 catalysts [11]. Reproduced by permission of
Imperial College Press, London.

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

sions, where zeolites and related microporous materials


have been demonstrated to be superior catalysts.
3.12.1. Gas to Liquids (GTL)/conversion of synthesis
gas to fuel
Fuels production directly from syngas (in former
times obtained from coal) has been reported by Fischer
and Tropsch in 1923 for the rst time [38], using an alkali-promoted iron catalyst. Fuels manufactured via the
FischerTropsch route reveal an excellent quality since
they consist mainly of linear parans and a-olens
and do not contain sulfur and aromatics. A Co-containing catalyst is applied for the production of heavy parafns via the FischerTropsch route starting with natural
gas, a technology developed by Shell and named the
Shell Middle Distillate Synthesis (SMDS) route
[39,40]. In addition, diesel fuel (or gasoline) is produced
by hydrocracking of the more or less sulfur and nitrogen-free wax obtained through the SMDS process using
noble metal containing zeolites. The more restricted fuel
specications currently introduced in order to reduce the
environmental impact of hazardous emissions represent
a driving force with respect to an increased use of fuels
prepared via the FischerTropsch route as a blending
component of the gasoline and diesel pools in the future
[39].
Besides the SMDS technology an alternative has been
presented by SASOL/Chevron termed as the SlurryPhase-Distillate process, again based on the Fischer
Tropsch route producing wax (using a Co-containing
catalyst) followed by a hydrocracking step in order to
get diesel or gasoline [41].
The methanol to gasoline (MTG) plant in New
Zealand has been combined with a methane steam
reforming unit for production of synthesis gas and a
methanol plant to produce gasoline from natural gas.
The process economics can be improved considerably
by a clever combination and close integration of the different steps. In the TIGAS process developed by Haldor
Topse AS for the manufacture of gasoline in a pilot
plant scale, the methanol synthesis and the MTG reactions are integratedwithout the separation of methanol as an intermediate product. A multi-functional
catalyst has been developed, however, these process
technologies do not usually apply catalysts based on
zeolites or related microporous materials [36]. Finally,
ExxonMobil has introduced the so-called Advanced
Gas Conversion for the 21st Century (AGC-21) technology, again based on the FischerTropsch route [131].
3.12.2. Methanol to hydrocarbons
Besides the direct route from synthesis gas to hydrocarbons (GTL-technology, based on the FischerTropsch route), a strong focus has been concentrated on
the indirect route via the production of methanol from
synthesis gas (mixture of carbon monoxide and hydro-

283

gen), which is made by steam reforming of natural gas


or gasication of coal, and the consecutive formation
of hydrocarbons. Of coarse, these technologies are alternatives to the chemical conversion of methane, either via
direct coupling, which is thermodynamically not favourable, or via oxidative coupling, a route not successfully
so far from an industrial point of view.
To begin with, the methanolhydrocarbons technology was primarily regarded as a powerful method to
convert coal into high-octane gasoline. This concept
has been expanded since, not only with respect to the
formation of other fuels, but also to chemicals in general. Of coarse, light olens are important components
for the petrochemical industry, and the demand of
high-quality gasoline is increasing as well. In fact, with
this new technology, one can make almost anything
out of coal or natural gas that can be made out of crude
oil.
Methanol is converted into an equilibrium mixture
of methanol, dimethylether and water, which can be
processed catalytically to either gasoline (methanol to
gasoline, MTG) or olens (methanol to olens, MTO),
depending on the catalyst and/or the process operation
conditions (see Fig. 26).
Although methanol itself is a potential motor fuel or
can be blended with gasoline, it would require large
investments to overcome the technical problems connected with it. The commercial MTG reaction runs at
temperatures around 400 C at a methanol pressure of
several bars and uses a ZSM-5 catalyst. These are the
optimal conditions for converting the olens that form
within the catalyst into parans and aromatics. However, at one point in the MTG reaction, the product
mixture consists of about 40% light olens. The importance of light olens as intermediates in the conversion
of methanol to gasoline was recognised early. Consequently, a number of attempts were made to selectively
form light olens from methanol, not only on mediumpore zeolites but also on small-pore zeolites, SAPO type
molecular sieves and over large-pore zeolites (however,
to a much lesser extent). If one interrupted the reaction
at the point of about 40% light olen formation, one
could harvest these C2C4 olens. By adjusting the reaction conditions (such as, for example, raising the temperature to 500 C) as well as the catalyst applied, one
can increase dramatically the olen yield. This discovery
led to the development of the MTO process, which generates mostly propene and butenes, with high-octane
gasoline as a by-product. However, the catalyst can be
modied in such a way that even more ethene is produced [132].
1973 marked the beginning of the energy crisis, and
new interest in synfuels and other chemicals favoured
the continuation of the methanolhydrocarbon research
[133,134]. Already the MTG and MTO processes represent a sort of chemical factory, to be brought on stream

284

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

Fig. 26. Methanol to hydrocarbons reaction path [132]. Reproduced by permission of Elsevier, Amsterdam.

as the technological and/or economic demands arise.


One can go a step further and convert the olens to
an entire spectrum of products, through another ZSM5 based process: Mobils olengasoline and distillate
process (MOGD), originally developed as a renery process, which works well coupled with the MTO process.
In the MOGD reaction, ZSM-5 oligomerises light olens, from either renery streams or MTO, into highermolecular-weight olens that fall into the gasoline,
distillate and lubricant range (see also Fig. 27) [135].
In 1979 the New Zealand government selected the
MTG process over the FischerTropsch (SASOL) process for converting natural gas from their extensive
Maui eld to gasoline. At that time, Mobils xed-bed
MTG process was unproven commercially, whereas
the SASOL technology was already commercialised
[135]. The New Zealand plant started to produce about
600 000 ton per year gasoline from April 1986, supplying
one-third of the nations gasoline demand [133]. The
gasoline production part of the factory was later closed
down, due to the price available for gasoline vs. the price

of methanol, however, the methanol production part is


still in operation.
In the scale-up to commercial operation of the MTG
plant two factors were of certain interest, catalyst deactivation and heat production, respectively (the MTG
reaction is highly exothermic). Despite the relatively
high stability and selectivity of the ZSM-5 catalyst
(due to the high silica content and the catalytic eectiveness based on the unique pore structure), deactivation
during the MTG operation is pronounced. The catalyst
is progressively coked and must be regenerated by calcination in air. The heat production in the MTG process
is high, leading to an adiabatic temperature rise of some
650 C [36,136,137].
The MTO technology seems now to be ready for
commercial use. The MTO process of Mobil has been
demonstrated in the same experimental 4 000 ton per
year plant at Wesseling (Germany) used to prove the
uid-bed MTG process, applying ZSM-5 as catalyst
[133]. The uid-bed technology provided all advantages
in terms of increased product yield, better quality and a

Fig. 27. Gasoline and distillate production via methanol and Mobils ZSM-5 technology [132]. Reproduced by permission of Elsevier, Amsterdam.

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

very ecient heat recovery. Depending on the world


market situation (price, demand, etc.) the uidised-bed
technology is available to produce liquid fuels via methanol [138].
UOP has, in co-operation with Norsk Hydro, announced in 1996 their SAPO-34 based MTO process
to be realised for construction of a 250 000 ton per year
plant using a natural gas feedstock for production of
ethene. A 0.5 ton per year demonstration unit operated
by Norsk Hydro has veried the olen yields and catalyst performance. SAPO-34 is extremely selective towards ethene and propene formation (about 80%) with
the exibility of altering the ratio between the two olens by varying the reactor conditions. The MTO process
can be designed for an ethene to propene ratio between
0.75 and 1.5at nearly complete methanol conversion.
The high selectivity to ethene gives SAPO-34 (which
has the chabazite (CHA) structure) a signicant advantage over other types of catalyst systems, like ZSM-5 or
SSZ-13 (synthetic aluminosilicate with CHA structure).
In addition, the SAPO-34 has a signicantly better stability due to a lower rate of coke formation than the
other catalytic systems with comparable acid site densities. The need to remove the high exothermic heat of the
MTO reaction as well as the need for frequent regeneration led to a uidised-bed reactor and regenerator
design. UOP and Norsk Hydro have commercially
manufactured the MTO catalyst (MTO-100TM), based
on SAPO-34, which has shown the type of attrition
resistance and stability suitable to handle multiple
regeneration steps and uidised-bed conditions [139,
140].
The discovery of the MTG reaction happened by
accident. One group at Mobil was trying to convert
methanol to other oxygen-containing compounds over
a ZSM-5 catalyst. Instead, they received unwanted
hydrocarbons. Somewhat later, another Mobil group,
working independently, was trying to alkylate isobutane
with methanol over ZSM-5 and identied a mixture of
parans and aromatics boiling in the gasoline range
all coming from methanol. Although the discovery of
MTG was accidental, it occurred due to a balanced effort in catalysis over many years. The MTO reaction
seems to benet from this development, although independent research has been performed since. The evolution of the methanolhydrocarbons technology, from
its discovery until its realisation on a demonstration
and/or commercial scale, has been accompanied by
extensive research related to the basic question of the
mechanism of formation of the initial CC bond. To address this topic in detail would be out of scope for this
paper, however, the interested reader can follow this discussion in the reviews mentioned in the list of Refs.
[132137].
Finally, an alternative to produce propene from
methanol has been introduced by Lurgi, the so-called

285

Methanol-to-Propene (MTP) process, applying an


H-ZSM-5 based catalyst from Sud-Chemie AG
[42,131].

4. Catalytic probe reactions with respect to the pore


architecture characterisation of zeolites and related
porous materials
4.1. General remarks
Dierent types of probe molecules have been used to
investigate various properties of zeolites and related
microporous and mesoporous materials, for example,
the pore architecture with respect to internal channel
dimensions, possible network connectivities, contributions of the external surface to the overall behaviour
of the materials, accessibility of pockets and caves in
connection with the structure of those compounds. Furthermore, probe molecules are used to evaluate the surface hydrophobicity and hydrophilicity as well as to
determine the acidity and basicity of porous systems.
These type of investigations are done either by physico-chemical studies of the interaction of probe molecules with the surface of the porous material (for
example by adsorption of the probe molecule and monitoring the interaction by applying spectroscopic, calorimetric, volumetric, gravimetric or other methods) or by
using probe molecules as model substrates in catalytic
reactions. Since the pores of zeolites and related microporous materials have dimensions comparable to those
of actual probe molecules, the phenomenon of shapeselectivity in catalytic reactions can be observed. However, in order to take advantage of this, the catalytic
reaction must take place at active sites on the internal
surface and not on the external one. Therefore, the
probe molecules applied are usually designed with respect to evaluate the inuence of both the internal and
external surface activities on the properties under investigation, like for example, a mixture of linear and
branched/bulky molecules. Branched and bulky probe
molecules are usually incapable to penetrate the micropores whereas linear molecules can.
Concerning the selection of suitable probe molecules
one should note that the crystallographic pore diameters
do not represent suitable threshold molecular diameters,
since molecules about 20% larger than this diameter can
be accommodated, especially at elevated temperatures.
Furthermore, the type of activity under investigation
(acid catalysed reaction, hydrogenation, oxidation etc.)
will inuence the choice of the probe molecules as well.
In addition, framework aluminium tends to reduce the
pore volume and to broaden the pore size distribution.
Finally, pre-adsorption of polar molecules like ammonia
or water can reduce the apparent pore size and, therefore, certain attention should be paid with respect to

286

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

the hydration state of the zeolite or related porous materials [141,142].


Catalytic probe reactions for studying the pore architecture of zeolites and related porous materials have
much in common with adsorption tests with the same
objective. Of course, adsorption takes place during these
catalytic tests as well. However, a shape-selective chemical reaction occurs in addition. The basic understanding
of shape-selective catalysis is, therefore, an important
item with respect to the design and appropriate interpretation of the results of catalytic test reactions.
This review deals with the topic of catalysis by zeolites. Consequently, the focus within this chapter will
be concentrated on catalytic reactions using probe
molecules. However, the interested reader will nd excellent reviews in the literature addressing the physicochemical studies of the interaction of probe molecules
with the surface of porous materials as well [141145].
A number of catalytic test reactions for the pore
architecture characterisation of zeolites and related porous materials have been introduced so far, and the
majority of the published data refer to Brnsted acid
catalysis. In the following sub-chapters these test reactions will be presented and discussed, including recent
advances within catalytic test reactions using probe
molecules.
4.2. Constraint index (CI)
The constraint index (CI) introduced by researchers
from the Mobil Oil Company more than 20 years ago,
has been the rst technique with respect to the characterisation of the relative pore width and shape-selective
properties of zeolites using a catalytic test reaction
[146]. This method is based on the competitive cracking
of an equimolar mixture of n-hexane and 3-methylpentane on acid zeolites. Originally, the test reaction was
designed to distinguish between small-, medium- and
large-pore zeolites, composed of 8-, 10- and 12-membered ring systems, respectively [147]. Based on the
shape-selective eect, the CI was dened as the ratio
of rst-order rate constants (k) of the cracking of n-hexane and 3-methyl-pentane:
Constraint index CI kn-hexane=k3-methylpentane
As long as the catalyst pores are suciently spacious,
branched alkanes are cracked at higher rates than their
linear isomers. In the literature, the following experimental conditions for the determination of the constraint index have been given: reaction temperature
between 290 and 510 C, LHSV between 0.1 and
1 h1, 10 vol.% of each reactant in He as carrier gas,
catalyst mass of 1 g, xed bed reactor at atmospheric pressure and an overall conversion of 1060%
[141,146].

A higher constraint index arises from the preferential


cracking of n-hexane compared with the branched isomer. The 3-methylpentane would be easier to crack in
the absence of steric hindrance. According to the Mobil
researchers, the constraint index can be used to classify
the molecular sieves into small-, medium- and large-pore
zeolites and related microporous materials [146]:
Small-pore (8-MR) systems
Medium-pore (10-MR) systems
Large-pore (12-MR) systems

12 < CI
1 < CI < 12
CI < 1

The attempt in the original paper by Frilette et al.


[146] was to provide a guideline for the determination
of small-, medium- and large-pore zeolites. However,
the discovery of new zeolite structure types has lead to
CI values which are misleading to incorrect conclusions
about the pore size and structure. Among those examples are zeolites with 14-membered ring systems and
as well as zeolites with
pore openings larger than 8 A
large internal cavities and pores composed of 8- or 9membered ring systems [147].
In conclusion, the introduction of the constraint
index represents the rst example of probing the pore
architecture of zeolites and related microporous materials by applying a catalytic test reaction. In spite of a
number of disadvantages, the constraint index has been
extensively used, and a number of literature data are
available [141,146,147].
4.3. Modied or rened constraint index
Whereas the constraint index acts as a test reaction
for mono-functional acidic molecular sieves, completely
dierent reaction mechanisms apply concerning test
reactions dedicated to bi-functional zeolites and related
microporous materials. The search for an appropriate
test reaction covering the use of bi-functional molecular
sieves has been concentrated on the isomerisation and
hydrocracking of long-chain n-alkanes. The common
feature of these reactions is the fact that they are performed using hydrogen, which is activated by the noble
metal component of the catalyst. Therefore, the isomerisation of n-decane at low conversions has been used to
dene the modied or rened constraint index CI* for
the characterisation of bi-functional zeolites [148]:
Modified or refined constraint index CI
Yield2-methylnonane =Yield5-methylnonane
The modied or rened constraint index CI* is now well
approved concerning the pore width characterisation of
medium-pore zeolites, that means 10-membered ring
molecular sieve systems. However, the nature and exact
origin of the shape-selective eects on which the index is
based is not completely understood yet. Since 2-methyl-

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

287

Fig. 29. Spaciousness index (SI) for dierent zeolites and related
microporous materials [141] (data taken from Ref. [155]). Reproduced
by permission of Wiley-VCH, Weinheim.
Fig. 28. Modied or rened constraint index (CI*) for dierent zeolites
[141] (data taken from Refs. [149151]). Reproduced by permission of
Wiley-VCH, Weinheim.

nonane is the kinetically preferred isomer in 10-membered ring zeolites, the amount of 2-methylnonane
formed from n-decane at low conversions increases relative to the other methylnonanes with decreasing pore
width of the zeolite.
Fig. 28 summarises CI* values taken from the literature [149151]. The CI* values for 10-membered ring
zeolites cover a broad range from about 3 to 15, which
represents a range where the CI* is quite suitable.
The only feature which the CI and CI* values have in
common is that their numerical values increase with
decreasing pore size of the zeolites under investigation.
On the other hand, the modied or rened constraint
index CI* is of little use concerning the pore width characterisation of 12-membered ring zeolites or related
microporous materials. However, there is another index
based on a dierent test reaction which is complementary to the CI* value, the so-called spaciousness index
(SI) for the characterisation of 12-membered zeolites
[141].
4.4. Spaciousness index
The mechanisms of hydrocracking and isomerisation
of cycloalkanes and n-alkanes are essentially identical.
However, dierent carbon number distributions have
been observed for the hydrocracking of n-alkanes and
cycloalkanes, respectively, both of them consisting of
10 carbon atoms. Weitkamp et al. registered during
hydrocracking of C10 cycloalkanes the formation of isobutane and methylcyclopentane almost exclusively in
the absence of spatial constraints. The carbocation
intermediates which govern the selectivity of hydrocracking of C10 cycloalkanes seem to be perfectly suited
for investigating the space availablecovering the entire
range of 12-membered ring zeolites and related microporous materials [141,152,153].
The spaciousness index (SI) is dened as the yield
ratio of isobutane and n-butane in the hydrocracked

products of a C10 cycloalkane (butylcyclohexane or pentylcyclopentane) [154,155]:


Spaciousness index SI Yieldisobutane =Yieldn-butane
SI values of dierent zeolites and related microporous
materials are given in Fig. 29. There is no doubt that
the spaciousness index is quite suitable for evaluation
of 12-membered ring microporous materials, extending
a broad range of SI values from about 3 to 20. On the
other hand, 10-membered ring zeolites have all together
a spaciousness index of about 1, indicating that the SI
system is not appropriate for the pore width evaluation
of medium-pore zeolites.
In conclusion, the spaciousness index is the method
of choice with respect to the pore width characterisation
of large-pore zeolites and related microporous materials
[141].
4.5. Disproportionation of ethylbenzene
As discussed in Section 3.10.2, the disproportionation
of ethylbenzene to benzene and diethylbenzenes has
been studied extensively [119,120]: With large pore zeolites (like Y zeolite), the reaction occurs via a hydride
transfer chain reaction through diphenylethanes as
intermediates. However, medium pore zeolites (like
ZSM-5) cannot accommodate this bulky intermediate,
and ethylbenzene disproportionation proceeds via a
dealkylationrealkylation path. To begin with, this reaction has been proposed in order to receive information
about the number of strong Brnsted acid sites in zeolites, however, later the reaction was found to be suitable for monitoring of the pore width as well [156].
Comparative experiments revealed that 12-membered
ring zeolites showed an induction period (followed by
no or limited deactivation) whereas 10-membered ring
systems did not, however, considerable deactivation
was observed in the case of medium-pore zeolites
[157,158]. In conclusion, the disproportionation of ethylbenzene allows a safe discrimination between largeand medium-pore zeolites and related microporous
materials, however, so far a clear ranking of these

288

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

Fig. 30. Isomerisation and disproportionation of m-xylene using acidic catalysts: Main reaction pathways [141]. Reproduced by permission of WileyVCH, Weinheim.

systems according to their eective pore width is not


possible [141]. Finally, the Catalysis Commission of
the International Zeolite Association (IZA) has recommended the disproportionation of ethylbenzene using
LaNaY zeolite as a standard reaction for acidity characterisation of acid zeolites [121].
4.6. Isomerisation and disproportionation of m-xylene
It is well known that m-xylene can perform both
isomerisation into o- and p-xylene as well as disproportionation into trimethylbenzene isomers and toluene
under acidic conditions (see Fig. 30) [159,160].
Gnep et al. suggested for the rst time the application
of m-xylene conversion for exploring the eective pore
width of zeolites more than two decades ago [161]. They
proposed several criteria with respect to the characterisation of zeolites and related microporous materials in
terms of pore architecture:
1. the relative rates of the formation of o- and p-xylene,
2. the rate ratio between isomerisation and disproportionation and
3. the distribution of the trimethylbenzenes as products
of the disproportionation.
The rst criterion is based on the observation that oand p-xylene are formed at about the same rates as long
as no shape-selective eect is in operation. However,
with decreasing pore width, the formation of the paraisomer is increasingly preferred in comparison to the
bulkier ortho-isomer (product shape selectivity).
The second criterion refers to the nding that isomerisation of m-xylene is more favoured compared to disproportionation in the case of decreasing pore width,
since suppression of the disproportionation is due to
the restricted transition state shape selectivity rather
than to mass transfer limitations.

Finally, the isomer distribution of the trimethylbenzenes reects the restricted transition state selectivity as
well, when discussing the hindered formation of 1,3,5trimethylbenzene in mordenite as compared to a larger
amount of this isomer using zeolite Y as catalyst.
The m-xylene test reaction has found certain interest
with respect to probing the pore architecture of zeolites
and related microporous materials, however, the application has been limited concerning the determination
of the pore width of those porous solids [141].

5. Conclusions and outlook


During the last 30 years the main focus on zeolites
and related microporous materials as industrial catalysts
has been concentrated on crude oil and natural gas
upgrading as well as petrochemical industrymainly
based on the acid properties of those porous materials.
To some minor extent, zeolites have been used as redox
catalysts as well. However, new demands and challenges
require the use of porous catalysts, besides the traditional applications, within new elds as well. The production of ne and special chemicals, drugs and other
compounds will in the future to a much larger extent
be performed utilising the catalytic shape selective properties of porous materials. Furthermore, the changes in
order to adapt the demands and constraints imposed
by the environmental legislation will, to a large extent,
be pursued by the application of porous materials
[162]. Porous materials in this context will not only
cover traditional zeolites and related microporous materials, but also include mesoporous materials, micro- and
mesoporous composite materials, metal organic open
framework materials, porous alumino silicates, organicinorganic hybrid materials and others [163].
However, for a number of important applications,
the pore size of the microporous materials are too small

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

for processing bulkier molecules. As a consequence, new


porous materials with larger pores have to be prepared.
With respect to the improvement of the catalytic properties of porous materials and taking into account the processing of bulkier molecules, the following challenges
have to be considered: Synthesis of ultra-large pore
zeolites or related microporous materials as well as delaminated systems. Furthermore, the preparation of
nano-crystalline zeolites or related microporous materials seems to be an interesting alternative as well, allowing large ratios of external to internal surfaces to be
achieved. Consequently, the reaction of bulky molecules
will take place at the external surface and/or at the pore
mouth of the zeolites. The preparation of micro- and
mesoporous composite systems would complete the
application of nano-crystalline zeolites in this context
quite well. In addition, delaminated zeolites show a very
large external surface with excellent accessibilities to active sites for bulky molecules. Finally, the synthesis of
new structures with ultra-large pores and three-dimensional framework systems of connected pores would be
the most direct way to expand the possibilities of zeolites
and related microporous materials within catalysis. A
number of eorts in this direction are in progress [164].
Catalysis will have to play a major role in overcoming
the technical challenges we will face in the futuresome
of them will require scientic and technological breakthroughsand without doubt, porous materials will
have a strong impact in relation to this development
[162].

Acknowledgments
The author is indebted to the organisers of the 14th
International Zeolite Conference Pre-Conference School
for the invitation to present this contribution. Thanks
are due to Andreas C. Moller for his assistance in connection with the technical preparation of the gures.
Finally, the author gratefully acknowledges nancial
support from SINTEF and funding from the European
Commission in connection with the TROCAT project
(contract no. G5RD-CT-2001-00520) and BIOCAT
project (contract no. ENK6-CT-2001-00510).
References
[1] J.A. Lercher, A. Jentys, in: F. Schuth, K.S.W. Sing, J. Weitkamp
(Eds.), Handbook of Porous Solids, Wiley-VCH, Weinheim,
2002, p. 1097.
[2] A. Dyer, An Introduction to Zeolite Molecular Sieves, John
Wiley, Chichester, 1988, p. 121.
[3] J.A. Rabo, G.C. Gajda, Catal. Rev. 31 (1989) 385.
[4] J.A. Martens, P.A. Jacobs, in: H. van Bekkum, E.M. Flanigen,
P.A. Jacobs, J.C. Jansen (Eds.), second ed., Studies in Surface
Science and Catalysis, vol. 137, Elsevier, Amsterdam, 2001, p.
633.

289

[5] G.A. Olah, P. von R. Schleyer, Carbonium Ions, Wiley, New


York, 1968, p. 342.
[6] S. Kotrel, H. Knozinger, B.C. Gates, Micropor. Mesopor.
Mater. 3536 (2000) 11.
[7] B.C. Gates, J.R. Katzer, G.C.A. Schuit, Chemistry of Catalytic
Processes, McGraw-Hill, New York, 1979, p. 1.
[8] J. Weitkamp, P.A. Jacobs, J.A. Martens, Appl. Catal. 8 (1983)
123.
[9] A. Raichle, Y. Traa, F. Fuder, M. Rupp, J. Weitkamp, Oil Gas
Eur. Mag. 1 (2003) 33.
[10] W.O. Haag, R.M. Dessau, in: Proceedings of the 8th International Congress on Catalysis, DECHEMA, Frankfurt, 1984, p.
305.
[11] A. Corma, A. Martinez, in: M. Guisnet, J.-P. Gilson (Eds.),
Zeolites for Cleaner Technologies, Imperial College Press,
London, 2002, p. 29.
[12] D.W. Werst, P. Han, S.C. Choure, E.I. Vinokur, L. Xu, A.D.
Trifunac, L.A. Eriksson, J. Phys. Chem. B 103 (1999) 9219.
[13] A. Corma, A. Martinez, Catal. Rev.-Sci. Eng. 35 (1993) 483.
[14] J. Weitkamp, Y. Traa, in: G. Ertl, H. Knozinger, J. Weitkamp
(Eds.), Handbook of Heterogeneous Catalysis, Wiley-VCH,
Weinheim, 1997, p. 2039.
[15] A. Corma, A. Martinez, C. Martinez, J. Catal. 146 (1994) 185.
[16] K.P. de Jong, C.M.A.M. Mesters, D.G.R. Peferoen, P.T.M. van
Brugge, C. de Groot, Chem. Ing. Sci. 51 (1996) 2053.
[17] G.S. Nivarthy, K. Seshan, J.A. Lercher, Micropor. Mesopor.
Mater. 22 (1998) 379.
[18] M.F. Simpson, J. Wei, S. Sundaresan, Ind. Eng. Chem. Res. 35
(1996) 3861.
[19] P.B. Weisz, V.J. Frilette, J. Phys. Chem. 64 (1960) 382.
[20] E.G. Derouane, Z. Gabelica, J. Catal. 65 (1980) 486.
[21] S.M. Csicsery, J. Catal. 23 (1971) 124.
[22] M. Guisnet, J.-P. Gilson, in: M. Guisnet, J.-P. Gilson (Eds.),
Zeolites for Cleaner Technologies, Imperial College Press,
London, 2002, p. 1.
[23] D.S. Santilli, T.V. Harris, S.I. Zones, Micropor. Mater. 1 (1993)
329.
[24] R.L. Gorring, J. Catal. 31 (1973) 13.
[25] J.A. Martens, R. Parton, L. Uytterhoeven, P.A. Jacobs, G.F.
Froment, Appl. Catal. 76 (1991) 95.
[26] J.A. Martens, W. Souverijns, W. Verrelst, R. Parton, G.F.
Froment, P.A. Jacobs, Angew. Chem. 34 (1995) 2528.
[27] W. Souverijns, J.A. Martens, G.F. Froment, P.A. Jacobs, J.
Catal. 174 (1998) 177.
[28] J.A.M. Arroyo, G.G. Martens, G.F. Froment, G.B. Marin,
P.A. Jacobs, J.A. Martens, Appl. Catal A: General 192 (2000) 9.
[29] A. Pfenninger, in: Proceedings Symposium Industrial Applications of Zeolites, October 2225, 2000, Brugge, Belgium, Techn.
Inst. VZW, p. 73.
[30] Ch. Marcilly, in: A. Galarneau, F. Di Renzo, F. Fajula, J.
Vedrine (Eds.), Zeolites and Mesoporous Materials at the Dawn
of the 21st Century, Studies in Surface Science and Catalysis, vol.
135, Elsevier, Amsterdam, 2001, p. 37.
[31] M.W. Schoonover, M.J. Cohn, Topics Catal. 13 (2000) 367.
[32] Ch. Marcilly, Topics Catal. 13 (2000) 357.
[33] A. Dyer, An Introduction to Zeolite Molecular Sieves, John
Wiley, Chichester, 1988, p. 127.
[34] G. Martino, P. Courty, C. Marcilly, in: G. Ertl, H. Knozinger,
J. Weitkamp (Eds.), Handbook of Heterogeneous Catalysis,
Wiley-VCH, Weinheim, 1997, p. 1801.
[35] R.H. Jensen, in: M. Guisnet, J.-P. Gilson (Eds.), Zeolites for
Cleaner Technologies, Imperial College Press, London, 2002, p.
75.
[36] I.E. Maxwell, W.H.J. Stork, in: H. van Bekkum, E.M. Flanigen,
P.A. Jacobs, J.C. Jansen (Eds.), Introduction to Zeolite Science
and Practice, second ed., Studies in Surface Science and
Catalysis, vol. 137, Elsevier, Amsterdam, 2001, p. 747.

290

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

[37] J. Weitkamp, A. Raichle, Y. Traa, M. Rupp, F. Fuder, Chem.


Commun. (2000) 1133.
[38] F. Fischer, H. Tropsch, Brennsto Chem. 4 (1923) 276.
[39] A. Corma, A. Martinez, in: F. Schuth, K.S.W. Sing, J. Weitkamp
(Eds.), Handbook of Porous Solids, Wiley-VCH, Weinheim,
2002, p. 2825.
[40] S.T. Sie, M.M.G. Senden, H.M.H. van Wechum, Catal. Today 8
(1991) 371.
[41] SASOL/Chevron slurry phase distillate, in: Proceedings 17th
World Petroleum Congress, Rio de Janeiro, Brazil, September
16, 2002, p. 90.
[42] W. Liebner, H. Koempel, H. Bach, in: Proceedings 17th World
Petroleum Congress, Rio de Janeiro, Brazil, September 16,
2002, p. 1.
[43] D.M. Brouwer, Rec. Trav. Chim. Bays-Bas 87 (1968) 1435.
[44] S.T. Sie, in: G. Ertl, H. Knozinger, J. Weitkamp (Eds.),
Handbook of Heterogeneous Catalysis, Wiley-VCH, Weinheim,
1997, p. 1998.
[45] J. Weitkamp, in: J.A. Ward (Ed.), Hydrocracking and Hydrotreating, ACS Symposium Series 20, American Chemical Society,
Washington, DC, 1975, p. 1.
[46] F. Schmidt, E. Kohler, in: M. Guisnet, J.-P. Gilson (Eds.),
Zeolites for Cleaner Technologies, Imperial College Press,
London, 2002, p. 153.
[47] J.R. Kiovsky, W.J. Goyette, T.M. Notermann, J. Catal. 52
(1978) 25.
[48] L.A. Pine, P.J. Maher, W.A. Watcher, J. Catal. 85 (1984) 466.
[49] B. Beagley, J. Dwyer, F.R. Fitch, R. Mann, J. Walters, J. Phys.
Chem. 88 (1984) 1744.
[50] E.G. Derouane, in: B. Imelik, C. Naccache, G. Coudurier, Y.
Ben Taarit, J.C. Vedrine (Eds.), Catalysis by Zeolites, Studies in
Surface Science and Catalysis, vol. 20, Elsevier, Amsterdam,
1985, p. 221.
[51] F. Alvarez, F.R. Ribeiro, G. Giannetto, F. Chevalier, G. Perot,
M. Guisnet, in: P.A. Jacobs, R.A. van Santen (Eds.), Zeolites:
Facts, Figures, Future, Studies in Surface Science and Catalysis,
vol. 49, Elsevier, Amsterdam, 1989, p. 1339.
[52] A. Chica, A. Corma, J. Catal. 187 (1999) 167.
[53] J. Miller, in: J. Weitkamp, H.G. Karge, H. Pfeifer, W. Holderich
(Eds.), Zeolites and Related Microporous Materials: State of the
Art 1994, Studies in Surface Science and Catalysis, vol. 84,
Elsevier, Amsterdam, 1994, p. 2319.
[54] G.H. Unzelman, C.J. Wolf, in: W.F. Bland, R.L. Davidson
(Eds.), Petroleum Processing Handbook, McGraw Hill, New
York, 1976, p. 3.
[55] H.A. Cheney, C.L. Raymond, Trans. Am. Inst. Chem. Eng. 42
(1946) 595.
[56] H.W. Grote, Oil Gas J. 56 (1958) 73.
[57] P.J. Kuchar, J.C. Bricker, M.E. Reno, R.S. Haizmann, Fuel
Proc. Techn. 35 (1993) 183.
[58] D.N. Nakahura, Hydrocarbon Process. 77 (1998) 15.
[59] J. Szabo, J. Perrotey, G. Szabo, J.C. Duchet, D. Cornet, J. Mol.
Catal. 67 (1991) 79.
[60] P. Grandvallet, K.P. de Jong, H.H. Mooiweer, A.G.T.G.
Kortbeek, B. Kraushaar-Czarnetzki, European Patent Application No. 0.501.577, 1992.
[61] H.H. Mooiweer, K.P. de Jong, B. Kraushaar-Czarnetzki,
W.H.J. Stork, S.C. Krutzen, in: J. Weitkamp, H.G. Karge, H.
Pfeifer, W. Holderich (Eds.), Zeolites and Related Microporous
Materials: State of the Art 1994, Studies in Surface Science and
Catalysis, vol. 84, Elsevier, Amsterdam, 1994, p. 2327.
[62] J. Weitkamp, S. Ernst, Proceedings 13th World Petroleum
Congress, vol. 3, Wiley, Chichester, 1992, p. 315.
[63] J. Weitkamp, in: B. Imelik, C. Naccache, T. Ben Taarit, J.
Vedrine, G. Coudurier, H. Praliaud (Eds.), Catalysis by Zeolites,
Studies in Surface Science and Catalysis, vol. 5, Elsevier,
Amsterdam, 1980, p. 65.

[64] J. Weitkamp, Compendium 78/79, Erdol Kohle-Erdgas-Petrochem., Supplementary volume, Industrieverlag von Hernhaussen, Leinfelden-Echterdingen, 1978, p. 525.
[65] A. Corma, A. Martinez, C. Martinez, Appl. Catal. A: General
134 (1996) 169.
[66] F. Cardona, N.S. Gnep, M. Guisnet, G. Szabo, P. Nascimento,
Appl. Catal. A 128 (1995) 243.
[67] G. Gardos, L. Pechy, A. Redey, C. Okonji, Hungarian J. Ind.
Chem. 8 (1980) 363.
[68] G. Gardos, A. Redey, M. Kovacs, J. Kristof, Hungarian J. Ind.
Chem. 11 (1983) 409.
[69] M. Stocker, H. Mostad, T. Rrvik, Catal. Lett. 28 (1994) 203.
[70] M. Stocker, H. Mostad, A. Karlsson, H. Junggreen, B. Hustad,
Catal. Lett. 40 (1996) 51.
[71] H. Mostad, M. Stocker, A. Karlsson, H. Junggreen, B. Hustad,
in: H. Chon, S.K. Ihm, Y.S. Uh (Eds.), Progress in Zeolites and
Microporous Materials, Studies in Surface Science and Catalysis,
vol. 105, Elsevier, Amsterdam, 1997, p. 1413.
[72] S. Unverricht, S. Ernst, J. Weitkamp, in: J. Weitkamp, H.G.
Karge, H. Pfeifer, W. Holderich (Eds.), Zeolites and Related
Microporous Materials: State of the Art 1994, Studies in Surface
Science and Catalysis, vol. 85, Elsevier, Amsterdam, 1994, p. 1693.
[73] A. Corma, V. Gomez, A. Martinez, Appl. Catal. A 119 (1994)
83.
[74] A. Corma, A. Martinez, P.A. Arroyo, J.L.F. Monteiro, E.F.
Sousa-Aguiar, Appl. Catal. A 142 (1996) 139.
[75] I. Kiricsi, C. Flego, G. Bellussi, Appl. Catal. A 126 (1995) 401.
[76] C. Flego, L. Galasso, I. Kiricsi, M.G. Clerici, in: B. Delmon,
G.F. Froment (Eds.), Catalyst Deactivation 1994, Studies in
Surface Science and Catalysis, vol. 88, Elsevier, Amsterdam,
1994, p. 585.
[77] Y.F. Chu, A.W. Chester, Zeolites 6 (1986) 195.
[78] J. Weitkamp, P.A. Jacobs, in: L. Guczi, F. Solymosi, P. Tetenyi
(Eds.), New Frontiers in Catalysis, Studies in Surface Science
and Catalysis, vol. 75, Elsevier, Amsterdam, 1993, p. 1735.
[79] A. Corma, A. Martinez, C. Martinez, Catal. Lett. 28 (1994) 187.
[80] J.R. Bernard, in: L.V. Rees (Ed.), Proceedings of the 5th
International Zeolite Conference, Heyden, London, 1980, p. 686.
[81] G.B. McVicker, J.L. Kao, J.J. Ziemiak, W.E. Gates, J.L.
Robbins, M.M.J. Treacy, S.B. Rice, T.H. Vanderspurt, V.R.
Cross, A.K. Chosh, J. Catal. 139 (1993) 48.
[82] D. Ostgard, L. Kustov, K.R. Poeppelmeier, W.M.H. Sachtler, J.
Catal. 133 (1992) 342.
[83] W. Zhang, P.G. Smirniotis, Appl. Catal. A: General 168 (1998)
113.
[84] A. Rainis, Int. Pat. Appl. WO 9,213,045, 1992.
[85] S.I. Zones, D.L. Holtermann, A. Rainis, Int. Pat. Appl. WO
9,111,501, 1991.
[86] R. von Ballmoos, D.H. Harris, J.S. Magee, in: G. Ertl, H.
Knozinger, J. Weitkamp (Eds.), Handbook of Heterogeneous
Catalysis, Wiley-VCH, Weinheim, 1997, p. 1955.
[87] E.T. Habib Jr., X. Zhao, G. Yaluris, W.C. Cheng, L.T. Boock,
J.-P. Gilson, in: M. Guisnet, J.-P. Gilson (Eds.), Zeolites
for Cleaner Technologies, Imperial College Press, London,
2002, p. 105.
[88] S.T Sie, in: J.C. Jansen, M. Stocker, H.G. Karge, J. Weitkamp
(Eds.), Advanced Zeolite Science and Applications, Studies in
Surface Science and Catalysis, vol. 85, Elsevier, Amsterdam,
1994, p. 587.
[89] A. Corma, V. Fornes, A. Martinez, F.V. Melo, O. Pallota, in: P.
Grobet, W.J. Mortier, E.F. Vansant, G. Schulz-Eklo (Eds.),
Innovation in Zeolite Materials Science, Studies in Surface
Science and Catalysis, vol. 37, Elsevier, Amsterdam, 1988, p. 495.
[90] F.G. Dwyer, T.F. Degnan, in: J.S. Magee, M.M. Mitchell Jr.
(Eds.), Shape Selectivity in Catalytic Cracking in FCC: Science
and Technology, Studies in Surface Science and Catalysis, vol.
76, Elsevier, Amsterdam, 1993, p. 499.

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292


[91] Z.T. Li, W.Y. Shi, R.N. Pan, F.K. Jiang, Prepr. Amer. Chem.
Soc., Div. Pet. Chem. 38 (1993) 581.
[92] A. Fu, D. Hunt, J.A. Bonilla, A. Batachari, Oil Gas J. 96 (1998)
49.
[93] J.-O. Barth, Erdol Erdgas Kohle 119 (2003) 86.
[94] A. Corma, V. Gonzalez-Alfaro, A.V. Orchilles, Appl. Catal. 129
(1995) 203.
[95] I.E. Maxwell, J.K. Minderhoud, W.H.J. Stork, J.A.R. van Veen,
in: G. Ertl, H. Knozinger, J. Weitkamp (Eds.), Handbook of
Heterogeneous Catalysis, Wiley-VCH, Weinheim, 1997, p. 2017.
[96] J.A.R. van Veen, in: M. Guisnet, J.-P. Gilson (Eds.), Zeolites
for Cleaner Technologies, Imperial College Press, London, 2002,
p. 131.
[97] J. Scherzer, A.J. Gruia, Hydrocracking Science and Technology,
Marcel Dekker, New York, 1996.
[98] J.L. Lemberton, M. Touzeyidio, M. Guisnet, Appl. Catal. 54
(1989) 101.
[99] J.L. Lemberton, M. Touzeyidio, M. Guisnet, Appl. Catal. A:
General 79 (1991) 115.
[100] A.P. Bolton, in: Zeolite Chemistry and Catalysis, ACS Monograph Series 117, Washington, DC, 1976, p. 714.
[101] C. Marcilly, J.P. Frank, in: B. Imelik, C. Naccache, T. Ben
Taarit, J. Vedrine, G. Coudurier, H. Praliaud (Eds.), Catalysis
by Zeolites, Studies in Surface Science and Catalysis, vol. 5,
Elsevier, Amsterdam, 1980, p. 93.
[102] J.W. Ward, in: D.L. Trimm, S. Akashah, M. Absi-Halabi, A.
Bishara (Eds.), Catalysis in Petroleum Rening 1989, Studies in
Surface Science and Catalysis, vol. 53, Elsevier, Amsterdam,
1990, p. 417.
[103] A. Hennico, A. Billon, P.-H. Bigeard, Hydrocarbon Technol.
Int. (1992) 19.
[104] C.T. Kresge, M.E. Leonowicz, W.J. Roth, J.C. Vashuli, J.S.
Beck, Nature 359 (1992) 710.
[105] A. Corma, A. Martinez, V. Martinez-Soria, J.B. Monton, J.
Catal. 153 (1995) 25.
[106] F. Schuth, in: A. Galarneau, F. Di Renzo, F. Fajula, J. Vedrine
(Eds.), Zeolites and Mesoporous Materials at the Dawn of the
21st Century, Studies in Surface Science and Catalysis, vol. 135,
Elsevier, Amsterdam, 2001, p. 1.
[107] A. Dyer, An Introduction to Zeolite Molecular Sieves, John
Wiley, Chichester, 1988, p. 130.
[108] N.Y. Chen, R.L. Gorring, H.R. Ireland, T.R. Stein, Oil Gas J.
75 (1977) 165.
[109] M. Daage, in: M. Guisnet, J.-P. Gilson (Eds.), Zeolites
for Cleaner Technologies, Imperial College Press, London,
2002, p. 167.
[110] D.V. Law, Institute of Petroleum Economics Rening Conference, London, UK, October 19, 1993.
[111] J.S. Beck, W.O Haag, in: G. Ertl, H. Knozinger, J. Weitkamp
(Eds.), Handbook of Heterogeneous Catalysis, Wiley-VCH,
Weinheim, 1997, p. 2123.
[112] J.S. Beck, A.B. Dandekar, T.F. Degnan, in: M. Guisnet, J.-P.
Gilson (Eds.), Zeolites for Cleaner Technologies, Imperial
College Press, London, 2002, p. 223.
[113] W.W. Kaeding, L.B. Young, C.T.-W. Chu, J. Catal. 89 (1984)
267.
[114] B.D. Flockhart, K.Y. Liew, R.C. Pink, J. Catal. 72 (1981)
314.
[115] T. Yashima, K. Sato, T. Hayasaka, N. Hara, J. Catal. 26 (1972)
303.
[116] J.A. Cusumano, ChemTech 22 (1991) 482.
[117] J.S. Beck, W.O Haag, in: G. Ertl, H. Knozinger, J. Weitkamp
(Eds.), Handbook of Heterogeneous Catalysis, Wiley-VCH,
Weinheim, 1997, p. 2136.
[118] F. Alario, M. Guisnet, in: M. Guisnet, J.-P. Gilson (Eds.),
Zeolites for Cleaner Technologies, Imperial College Press,
London, 2002, p. 189.

291

[119] M. Weihe, M. Hunger, M. Breuninger, H.G. Karge, J. Weitkamp, J. Catal. 198 (2001) 256.
[120] N. Arsenova-Hartel, H. Bludau, W.O. Haag, H.G. Karge,
Micropor. Mesopor. Mater. 35 (2000) 113.
[121] D.E. De Vos, S. Ernst, C. Perego, C.T. OConnor, M. Stocker,
Micropor. Mesopor. Mater. 56 (2002) 185.
[122] C. Ringelhan, V. Kurth, G. Burgfels, J.G. Neumayr, W. Seuert,
J. Klose, Erdol Erdgas Kohle 118 (2002) 88.
[123] A. Raichle, T. Traa, J. Weitkamp, Chem. Ing. Tech. 73 (2001)
947.
[124] J. Weitkamp, A. Raichle, Y. Traa, Appl. Catal A: General 222
(2001) 277.
[125] A. Raichle, Y. Traa, J. Weitkamp, Erdol Erdgas Kohle 118
(2002) 83.
[126] J. Weitkamp, A. Raichle, Y. Traa, M. Rupp, F. Fuder, Chem.
Commun. (2000) 403.
[127] H. Bischof, W. Dohler, F. Fuder, J. Laege, Erdol Erdgas Kohle
118 (2002) 79.
[128] C.T. OConnor, in: G. Ertl, H. Knozinger, J. Weitkamp (Eds.),
Handbook of Heterogeneous Catalysis, Wiley-VCH, Weinheim,
1997, p. 2069.
[129] A. Hagen, F. Rossner, Catal. Rev.-Sci. Eng. 42 (2000) 403.
[130] K.-H. Steinberg, U. Mroczek, F. Rossner, Appl. Catal. 66 (1990)
37.
[131] S. Ernst, Erdol Erdgas Kohle 119 (2003) 91.
[132] M. Stocker, Micropor. Mesopor. Mater. 29 (1999) 3.
[133] C.D. Chang, J.A. Silvestri, ChemTech 10 (1987) 624.
[134] C.D. Chang, in: G. Ertl, H. Knozinger, J. Weitkamp (Eds.),
Handbook of Heterogeneous Catalysis, Wiley-VCH, Weinheim,
1997, p. 1894.
[135] S.L. Meisel, ChemTech 1 (1988) 32.
[136] C.D. Chang, Catal. Rev.-Sci. Eng. 25 (1983) 1.
[137] C.D. Chang, Catal. Rev.-Sci. Eng. 26 (1983) 323.
[138] H.R. Grimmer, N. Thiagarajan, E. Nitschke, in: D.M. Bibby,
C.D. Chang, R.F. Howe, S. Yurchak (Eds.), Methane Conversion, Studies in Surface Science and Catalysis, vol. 36, Elsevier,
Amsterdam, 1988, p. 273.
[139] B.V. Vora, T.L. Marker, P.T. Barger, H.R. Nilsen, S. Kvisle, T.
Fuglerud, in: M. de Pontes, R.L. Espinoza, C.P. Nicolaides,
J.M. Scholz, M.S. Scurrell (Eds.), Natural Gas Conversion IV,
Studies in Surface Science and Catalysis, vol. 107, Elsevier,
Amsterdam, 1997, p. 87.
[140] P. Barger, in: M. Guisnet, J.-P. Gilson (Eds.), Zeolites for
Cleaner Technologies, Imperial College Press, London, 2002, p.
239.
[141] Y. Traa, J. Weitkamp, in: F. Schuth, K.S.W. Sing, J. Weitkamp
(Eds.), Handbook of Porous Solids, Wiley-VCH, Weinheim,
2002, p. 1015.
[142] P.J. Kunkeler, R.S. Downing, H. van Bekkum, in: H. van
Bekkum, E.M. Flanigen, P.A. Jacobs, J.C. Jansen (Eds.),
Introduction to Zeolite Science and Practice, second ed., Studies
in Surface Science and Catalysis, vol. 137, Elsevier, Amsterdam,
2001, p. 987.
[143] J. Weitkamp, S. Ernst, Catal. Today 19 (1994) 107.
[144] L. Forni, Catal. Today 41 (1998) 221.
[145] J. Weitkamp, S. Ernst, Catal. Today 3 (1988) 451.
[146] V.J. Frilette, W.O. Haag, R.M. Lago, J. Catal. 67 (1981)
218.
[147] S.I. Zones, T.V. Harris, Micropor. Mesopor. Mater. 3536
(2000) 31.
[148] J.A. Martens, M. Tielen, P.A. Jacobs, J. Weitkamp, Zeolites 4
(1984) 98.
[149] A. Corma, C. Corell, F. Llopis, A. Martinez, J. Perez-Pariente,
Appl. Catal. A: General 115 (1994) 121.
[150] P.A. Jacobs, J.A. Martens, Pure Appl. Chem. 58 (1986) 1329.
[151] W. Souverijns, L. Rombouts, J.A. Martens, P.A. Jacobs,
Micropor. Mater. 4 (1995) 123.

292

M. Stocker / Microporous and Mesoporous Materials 82 (2005) 257292

[152] J. Weitkamp, S. Ernst, H.G. Karge, Erdol Kohle-ErdgasPetrochem. 37 (1984) 457.


[153] S. Ernst, J. Weitkamp, in: Proceedings International Symposium
on Zeolite Catalysis, Siofok, Hungary, May 1316, 1985, Acta
Phys. Chem. (1985) 457.
[154] J. Weitkamp, S. Ernst, R. Kumar, Appl. Catal. 27 (1986)
207.
[155] J. Weitkamp, S. Ernst, C.Y. Chen, in: P.A. Jacobs, R.A. van
Santen (Eds.), Zeolites: Facts, Figures, Future, Studies in Surface
Science and Catalysis, vol. 49, Elsevier, Amsterdam, 1989, p.
1115.
[156] H.G. Karge, J. Ladebeck, Z. Sarbak, K. Hatada, Zeolites 2
(1982) 94.

[157] J. Weitkamp, S. Ernst, P.A. Jacobs, H.G. Karge, Erdol KohleErdgas-Petrochem. 39 (1986) 13.
[158] H.G. Karge, Y. Wada, J. Weitkamp, S. Ernst, U. Girrbach,
H.K. Bayer, in: S. Kaliaguine, A. Mahay (Eds.), Catalysis on the
Energy Scene, Studies in Surface Science and Catalysis, vol. 19,
Elsevier, Amsterdam, 1984, p. 101.
[159] S.M. Csicsery, J. Org. Chem. 34 (1969) 3338.
[160] A. Corma, E. Sastre, J. Catal. 129 (1991) 177.
[161] N.S. Gnep, J. Tejada, M. Guisnet, Bull. Soc. Chim. Fr. 1 (1982)
5.
[162] C. Marcilly, J. Catal. 216 (2003) 47.
[163] Y. Liu, Th.J. Pinnavaia, J. Amer. Chem. Soc. 125 (2003) 2376.
[164] A. Corma, J. Catal. 216 (2003) 298.

Anda mungkin juga menyukai