You are on page 1of 46

Microporous and Mesoporous Materials 29 (1999) 3–48

Review

Methanol-to-hydrocarbons: catalytic materials and their behavior1


Michael Stöcker
SINTEF Applied Chemistry, Department of Hydrocarbon Process Chemistry, P.O.Box 124 Blindern,
N-0314 Oslo, Norway
Received 7 November 1997; accepted 30 January 1998

Abstract

The literature related to methanol-to-hydrocarbons (MTHC ) technology over the past two decades has been
reviewed, covering mainly the methanol-to-olefin (MTO) and methanol-to-gasoline (MTG) reactions. The work
before around 1990 is briefly addressed, and the interested reader’s attention is drawn to review papers published
during the late 1980s summarizing the early work related to the methanol-to-hydrocarbons technology. This review
focuses mainly on the chemistry and mechanism of those reactions including the catalysts involved and their behavior
due to crystal size, pore architecture, acidity and reaction conditions, covering the time since around 1990. In a
second review authored by F. Keil, the process related items of methanol-to-hydrocarbons technology will be
summarized and discussed in the light of kinetic and reaction technology aspects. © 1999 Elsevier Science B.V. All
rights reserved.

Keywords: Methanol-to-hydrocarbons; MTO (methanol-to-olefin); MTG (methanol-to-gasoline); Heterogeneous


catalysis; Review

1. Introduction ogy, one can make almost anything out of coal or


natural gas that can be made out of crude oil.
To begin with, the methanol-to-hydrocarbons Methanol is made from synthesis gas (a mixture
technology was primarily regarded as a powerful of carbon monoxide and hydrogen) which is
method to convert coal into high-octane gasoline. formed by steam reforming of natural gas or
This concept has been expanded since, not only gasification of coal. The methanol is then con-
with respect to the formation of other fuels, but verted to an equilibrium mixture of methanol,
also to chemicals in general. Of course, light olefins dimethyl ether and water, which can be processed
are important components for the petrochemical catalytically to either gasoline (methanol-to-gaso-
industry, and the demand for high-quality gasoline line, MTG) or olefins (methanol-to-olefin, MTO),
is increasing as well. In fact, with this new technol- depending on the catalyst and/or the process oper-
ation conditions. Although methanol itself is a
potential motor fuel or can be blended with gaso-
1Dedicated to my wife Wencke Ophaug. line, it would require large investments to over-

1387-1811/99/$ – see front matter © 1999 Elsevier Science B.V. All rights reserved.
PII: S1 3 8 7 -1 8 1 1 ( 9 8 ) 0 0 31 9 - 9
4 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

come the technical problems connected with the the MTG process over the Fischer–Tropsch
direct use of methanol as a motor fuel. The com- (SASOL process) for converting natural gas from
mercial MTG reaction runs at temperatures their extensive Maui field to gasoline. At that time,
around 400°C and a methanol partial pressure of Mobil’s fixed-bed MTG process was unproven
several bars and uses a ZSM-5 catalyst. These are commercially, whereas the SASOL technology was
the optimal conditions for converting the olefins already commercialized [1]. The New Zealand
that form within the catalyst into paraffins and plant started to produce about 600 000
aromatics. However, at one point in the MTG ton year−1 gasoline from April 1986, supplying
reaction, the product mixture consists of about one-third of the nation’s gasoline demand [2].
40% light olefins. The importance of light olefins The gasoline production part of the factory was
as intermediates in the conversion of methanol to later closed down, due to the price available for
gasoline was recognized early. Consequently, a gasoline versus the price of methanol, however,
number of attempts were made to selectively form the methanol production part is still in
light olefins from methanol, not only on medium- operation.
pore zeolites but also on small-pore zeolites, SAPO The MTO technology seems now to be ready
type molecular sieves and large-pore zeolites (to a for commercial use. The MTO process of Mobil
much lesser extent). If one interrupted the reaction has been demonstrated in the same experimental
at the point of about 40% light olefin formation, 4000 ton year−1 plant at Wesseling (Germany)
one could harvest these C –C olefins. By adjusting used to prove the fluid-bed MTG process, applying
2 4 ZSM-5 as catalyst [2]. In cooperation with Norsk
the reaction conditions (such as, for example,
raising the temperature to 500°C ) as well as the Hydro, UOP announced in 1996 that their
SAPO-34 based MTO process was to be realized
catalyst applied, one can increase dramatically the
in the construction of a 250 000 ton year−1 plant
olefin yield. This discovery led to the development
using a natural gas feedstock for production of
of the MTO process, which generates mostly pro-
ethylene. A 0.5 ton year−1 demonstration unit
pylene and butylene, with high-octane gasoline as
operated by Norsk Hydro has verified the olefin
a byproduct. However, the catalyst can be modified
yields and catalyst performance. SAPO-34 is
in such a way that even more ethylene is pro-
extremely selective towards ethylene and propylene
duced [1].
formation with the flexibility of altering the ratio
1973 marked the beginning of the energy crisis,
between the two olefins by varying the reactor
and new interest in synfuels and other chemicals conditions [3].
favored the continuation of the methanol-to- The discovery of the MTG reaction happened
hydrocarbon research [2]. Already the MTG and by accident. One group at Mobil was trying to
MTO processes represent a sort of chemical convert methanol to other oxygen-containing com-
factory, to be brought on stream as the technologi- pounds over a ZSM-5 catalyst. Instead, they
cal and/or economic demands arise. One can go received unwanted hydrocarbons. Somewhat later,
one step further and convert the olefins to an another Mobil group, working independently, was
entire spectrum of products, through another trying to alkylate isobutane with methanol over
ZSM-5 based process: Mobil’s olefin-to-gasoline ZSM-5 and identified a mixture of paraffins and
and distillate process (MOGD), originally devel- aromatics boiling in the gasoline range – all coming
oped as a refinery process, which works well from methanol. Although the discovery of MTG
coupled with the MTO process. In the MOGD was accidental, it occurred due to a balanced effort
reaction, ZSM-5 oligomerizes light olefins, in catalysis over many years. The MTO reaction
from either refinery streams or MTO, into higher- seems to benefit from this development, although
molecular-weight olefins that fall into the gasoline, independent research has been performed since.
distillate and lubricant range (see also Fig. 1) The evolution of the methanol-to-hydrocarbons
[1]. (MTHC ) technology, from its discovery until its
In 1979 the New Zealand government selected realization on a demonstration and/or commercial
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 5

Fig. 1. Gasoline and distillate production via methanol and Mobil’s ZSM-5 technology (reproduced by permission of Elsevier Science
B.V., Amsterdam).

scale, has been accompanied by extensive research Methanol is first dehydrated to dimethyl ether
related to the basic question of the mechanism of (DME ). The equilibrium mixture formed, con-
formation of the initial C–C bond [1,2]. This sisting of methanol, dimethyl ether and water, is
review tries to highlight the milestones of this then converted to light olefins. In the last step of
development during the past decade. this scheme, the light olefins react to form paraffins,
A number of reviews exist dealing with the aromatics, naphthenes and higher olefins by
methanol-to-hydrocarbons technology, mainly hydrogen transfer, alkylation and polycondensa-
covering the MTG reaction, and the summaries tion (see Fig. 2).
by Chang and others should be particularly men- There is general consensus that the intermediate
tioned here [4–11]. I recommend the interested in the dehydration of methanol to dimethyl ether
reader to consult these reviews for the R&D work [step 1 in Eq. (1)] over solid acid catalysts is a
performed during the 1980s and earlier. This protonated surface methoxyl, which is subject to
review will focus on the MTO reaction, however, a nucleophilic attack by methanol [12]. The subse-
recent results covering the MTG reaction will be quent conversion of light olefins to paraffins, aro-
addressed as well. matics, naphthenes and higher olefins [step 3 in
Eq. (1)], which proceeds via classical carbenium
ion mechanisms with concurrent hydrogen
2. The chemistry of methanol-to-hydrocarbons: transfer, is well known from hydrocarbon chemis-
mechanistic and kinetic considerations try in acid media [11]. However, the second step
in Eq. (1), which represents the initial C–C bond
The main reaction steps of the methanol conver- formation from the C reactants, has been the
sion to hydrocarbons can be summarized as fol- 1
topic of an extensive discussion throughout the
lows:

(1)
6 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

Fig. 2. Methanol-to-hydrocarbons reaction path (reproduced by permission of Elsevier Science B.V., Amsterdam).

years. A large number of papers is available pre- ion, rather than an intramolecular Stevens
senting more than 20 possible mechanistic propos- rearrangement [12,17].
als for the formation of the first C–C bond. The Much research has been concentrated on the
most discussed and/or relevant of these mecha- questions about the existence of the oxonium
nisms can be broadly classified and briefly summa- ylides, their ability to rearrange according to
rized as follows [11–13]: Stevens and the zeolite’s ability to abstract a
The oxonium ylide mechanism has without doubt proton from the oxonium ions to form the desired
received strong attention. Van den Berg et al. [14] ylides. Again, Olah and coworkers performed a
postulated that dimethyl ether interacts with a number of experiments in an endeavor to answer
Brønsted acid site of the solid catalyst to form a these questions by attempting the synthesis of the
dimethyl oxonium ion, which reacts further with ylide directly via three different approaches: (1)
another dimethyl ether to form a trimethyl oxo- reaction of photochemically formed methyl car-
nium ion. This trimethyl oxonium ion is subse- bene with dimethyl ether; (2) deprotonation of the
quently deprotonated by a basic site to form a trimethyloxonium ion; and (3) desilylation of the
surface associated dimethyl oxonium methyl ylide dimethyl [(trimethylsilyl ) methyl ] oxonium ion
species [11,12]. The next step is either an intramo- [17,18]. No ylide species were isolated in these
lecular Stevens rearrangement [14], leading to the experiments, however, their existence was inferred
formation of methylethyl ether, or an intermolecu- from the product isotope distributions by using
lar methylation [15], leading to the formation of 13C or 2H labelled compounds. The observed iso-
the ethyldimethyl oxonium ion. In both cases tope distributions of methylethyl ether appeared
ethylene is formed via b-elimination (see Scheme 1) to rule out the Stevens rearrangement [12].
[11]. A bifunctional acid–base catalyzed condensa- Furthermore, Olah et al. investigated the photo-
tion was suggested by Olah [16 ]. Methanol and chemical reaction of diazomethane with dialkyl
dimethyl ether form a trimethyl oxonium ion ethers, using labelled and unlabelled CH N [18].
2 2
(using for example WO on alumina), which is Based on the analysis of the product isotope
3
deprotonated to dimethyl oxonium methyl ylide. distribution, they proposed the transmethylation
However, Olah concluded (based on labelling process yielding dimethyl ether via the formation
experiments) that the ylide undergoes bimolecular of methylethyloxonium methyl ylide, which either
methylation to form the ethyldimethyl oxonium decomposes to ethylene and dimethyl ether or is
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 7

Scheme 1

protonated by methanol (or water) impurities to are involved in the C–C bond formation, their is
give an oxonium salt, which then undergoes rapid still the question about the zeolite’s ability to
methyl transfer to form dimethyl ether [18]. These abstract a proton from the oxonium ions to form
observations were taken as an evidence against the the desired ylides. Hunter and Hutchings [21,22]
Stevens rearrangement as the predominant path- investigated this topic closely, concluding with the
way [12]. According to Huisgen, the reaction of suggestion that a surface-bound methoxyl is initially
diazomethane with dialkyl ethers proceeds via an formed, which is deprotonated to a surface-bound
initial ylide formation, followed by a Stevens ylide. This ylide is isoelectronic with surface-bound
rearrangement [19]. However, this was questioned carbene [23,24], which in turn engages in a C–C
by Franzen and Fikentscher [20], who concluded bond formation by C–H attack. However, the
that the formation of methylethyl ether and ethy- nature of the basic site for deprotonation was not
lene could be rationalized by b-elimination from a specified [12]. Another group investigated the ques-
methyleneoxonium ylide intermediate. tion of basicity by preparing the trimethyloxonium
A lot of suitable information was obtained by ion bound to ZSM-5 and following its decomposi-
running those experiments, however, they did not tion by means of 13C MAS NMR spectroscopy.
give a definitive answer with respect to the ylide Upon warming to room temperature, the sample
question as a result of their shortcomings. slowly decomposed to methyl-ZSM-5 and phy-
Nevertheless, if those oxonium ylides do exist and sisorbed dimethyl ether, confirming that the conju-
8 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

gate base of H-ZSM-5 is nucleophilic rather than The carbene mechanism involves the a-elimination
basic towards the methylating agent [12,25]. Similar of water from methanol followed by either poly-
results (formation of methyl-ZSM-5) were moni- merization of the resultant carbene to olefins or
tored by Forester et al. [26 ] applying in situ Fourier by concurrent sp3 insertion of the carbene into
transform infrared (FTIR) spectroscopy. methanol or dimethyl ether [36–38]. The forma-
Chao et al. studied the conversion of methanol tion of the carbene by the cooperative action of
to hydrocarbons over ZSM-5 with mixtures of acid and basic sites in mordenite was suggested by
labelled and unlabelled methanol and dimethyl Swabb and Gates [38] and can be summarized as
ether at elevated temperatures. Their results indi- follows:
cate that ethylene was the primary hydrocarbon
product through bimolecular reactions of dimethyl [Zeo−O−÷H–CH –OHH–O–Zeo]
2
ether with or without methanol on the catalysts H O+:CH (2)
surface. The C–C bond was formed following an 2 2
acidic mechanism via oxonium ions with intermo- According to Salvador and Kladnig [39], carbenes
lecular transfer of H and D. An isotope distribu- are generated through decomposition of surface
tion analysis of the products showed that methoxyls formed originally upon chemisorption
propylene was most likely produced by the reaction of methanol on the zeolite (zeolite Y ) [12].
of ethylene and C species from methanol adsorbed Venuto and Landis [37] addressed for the first
1
on the ZSM-5 surface [27]. time the subject of mechanism of methanol conver-
In addition, a number of other investigators sion to hydrocarbons over zeolites (zeolite X ),
favor the ‘oxonium proposal’ as the most likely whereas Chang and Silvestri [36 ] reported the
mechanism as a result of the linear dependence of conversion of methanol to hydrocarbons over
hydrocarbon formation activity on zeolite ZSM-5, favoring the carbene mechanism for the
Brønsted acidity [28–33]. initial C–C bond formation as well. Later, the
In summary, the oxonium ylide mechanism possibility of a sequential mechanism was consid-
involves the formation of a surface-bound interme- ered by Chang [40].
diate as the initial reaction step. The zeolitic Up to now, experimental evidence with respect
surface OH-group is methylated to form the to the intermediate carbene formation has been
methyloxonium intermediate, which gives rise to a indirect, for example by trapping the reactive C
1
surface-bound methylene-oxoniumylide due to intermediate from 13C labelled methanol decompo-
deprotonation. The surface-bound methylene oxo- sition over ZSM-5 using unlabelled propane. The
niumylide is isoelectronic with a surface-associated analysis of the product isomer and isotope distribu-
carbene, as outlined in Scheme 2 [34,35]. Further tions gave indications that the reactive C interme-
1
methylation results in the initial C–C bond forma- diate was carbenic and the mode of attack on
tion [11]. propane was sp3 insertion into H–C bonds [41].

Scheme 2
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 9

Such carbene insertions are indiscriminate in selec- of singlet methylene into trimethyloxonium tetra-
tivity [42], resulting in a statistical distribution of fluoroborate. On the other hand, the alkylation of
the butane isomers (i:n=1:3). Methylation by a diazomethane by alkyl cations is known.
carbocationic C intermediate would, by contrast, Consequently, Olah et al. suggested an alternate
1
result in high isobutane selectivity [43]. However, interpretation where a stabilized ‘incipient methyl
this interpretation was questioned by van Hooff cation’ on the acid site reacted with diazomethane
[44] who suggested that low i/n-butane ratios could to form ethylene. The carbene mechanism was
be explained by hydride transfer between propane further debated by Olah et al. [47] based on
and butyl cations. If hydride transfer were slower thermochemical arguments, however, calculated
to t-butyl cation than to the secondary cation, low stabilization energies in the range of 285–
i/n-butane ratios would result. Hydride transfer 314 kJ mol−1 showed that surface-stabilized diazo-
would also activate propane for C–C scission, methane is not an unreasonable intermediate for
leading to doubly labelled butanes [12]. Chang hydrocarbon formation from methanol [12,48].
and Chu [45] replied by showing that once propane Nováková et al. [49] assumed the formation of
enters into the main reaction path, the i/n-butane the first C–C bond in ethylene and propylene
ratios will be constant due to complete isotope during methanol transformation over zeolites to
scrambling, and independent of the degree of 13C occur via the reaction of gaseous methanol (or
substitution, contrary to the experimental observa- dimethyl ether) with C species whose C–H bonds
1
tion. Van Hooff ’s proposal also failed to account are strongly weakened. These species are probably
for the high i/n ratios observed in the reaction of polarized methoxyls or protonated formaldehyde
methanol alone [12]. which could also yield species of carbene-like char-
Evidence in favor of carbene intermediacy in acter. The authors observed, that on H-ZSM-5,
the conversion of methanol to gasoline over the lack of methanol in the gas phase led to a
H-ZSM-5 was reported by Dass et al. [46 ]. The disproportionation of methoxyls to methane and
authors observed a decrease in the abundance ratio formaldehyde. This was not observed on dealumi-
of isobutane to that of n-butane, observed on nated H–Y, probably due to the supply of metha-
addition of propane to the feed. The lowering of nol by the decomposed surface species [49].
the i/n-butane ratio may be accounted for in part Hutchings et al. [50] contributed with a number
by conversion of propane yielding butanes, pre- of papers dealing with the mechanistic aspects of
dominantly n-butane. An additional effect is the the methanol to hydrocarbons reaction, among
raise of the i/n-butane ratio on adding helium as them the reaction of methanol with added
diluent in the control sample. At least the greater hydrogen using WO on alumina and H-ZSM-5
3
part of the i/n-butane ratio lowering effect may be catalysts which did not significantly alter the pro-
attributed to these two causes. duct distribution from that of normal methanol
Lee and Wu [23] conducted a model study using conversion. The authors considered their results
diazomethane, which was reacted over various as a clear evidence against the involvement of a
catalysts including ZSM-5, with the aim of observ- gas phase carbene intermediate. In addition, the
ing a possible surface-stabilization of carbene. behavior of the methoxylmethyl radical in the gas
Although olefins were formed even with ‘inert’ phase has been studied and the results demonstrate
materials, hydrocarbon yields were seen to increase that this species is not involved either as an inter-
substantially in the presence of acidic surfaces. mediate in initial C–C bond formation in the
They proposed the formation of a methyloxonium methanol to hydrocarbons reaction [50].
species via interaction with an acid site. This Whereas the oxonium ylide mechanism essen-
species reacts with excess diazomethane, forming tially involves the formation of a surface-bound
an ethyloxonium species by C–H insertion, and intermediate as the initial reaction step, the carbene
final b-elimination affords ethylene [12]. The inter- mechanism involves only surface associated inter-
pretation of Lee and Wu was challenged by Olah mediates [11], see Scheme 2.
et al. [47] who were not able to induce the insertion Other authors/groups favored the carbocationic
10 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

mechanism [50–55]. According to Ono and Mori


[51] surface methoxyls may function as free methyl
cations, which adds to the C–H bond of dimethyl
ether to form a pentavalent carbonium transition
state. Abstraction of a proton completes the reac-
tion, analogous to the proposed ‘superacid mecha-
nism’ by Olah et al. [43]. However, one may argue
whether the C–H bond of methanol or dimethyl
Scheme 3
ether is sufficiently nucleophilic to undergo substi-
tution as proposed by Ono and Mori. This item
was investigated by Smith and Futrell [56 ] who this product has never been observed in the
reacted methylcarbenium ion with methanol. dimethyl ether reaction over zeolites [12].
Hydride abstraction to form methane and The free radical pathway has been discussed by
CH OH+ was the superior reaction (85–90%), several other authors concluding with experimental
2
however, proton transfer from the methyl cation evidence against a radical mechanism, like Hunter
to methanol (forming methylene) happened to a et al. [61] who studied the behavior of the
certain extent (7%). C species were not observed. methoxymethyl radical in the gas phase conversion
2
These results give rise to some doubt about the of methanol to hydrocarbons over ZSM-5. The
suggested mechanism, attacking directly the C–H same authors ended up with the same conclusion
bond of the methoxyl group by a methyl cation when investigating the reactions of dimethyl ether
[12]. and methanol over H-ZSM-5 with added NO and
As an alternative for the formation of methylox- O [62]. In continuation of this work, Hutchings
2
onium intermediates, the free radical mechanism compared the inhibition of the methanol conver-
has been introduced [57]. The participation of free sion to hydrocarbons over H-ZSM-5 induced by
radicals in the conversion of methanol to hydro- co-feeding NO with the inhibition by NH and
3
carbons over natural mordenite has been suggested NH OH [63]. It was considered that NO does not
2
by Zatorski and Krzyzanowski [58], however, act as a radical scavenger during this reaction, nor
without experimental evidence documented in their does NO co-feeding lead to the generation of
report. Later, Clarke et al. [59] recognized that NH as a catalyst poison. It was observed that
3
dimethyl ether could be the source of methylradi- inhibition induced by NH OH closely simulates
2
cals when they identified free radicals in the reac- that observed by NO co-feeding [63].
tion of dimethyl ether over H-ZSM-5 monitored Free radical signals have been detected during
by ESR spectroscopy. Clarke et al. proposed that ESR measurements carried out on the conversion
radicals are formed initially by interaction of of methanol to hydrocarbons over H-ZSM-5 in a
dimethyl ether with paramagnetic centers (solid- flow reactor at temperatures between 200°C and
state defects) in the zeolite and that C–C bond 280°C [64]. However, the signals were quenched
formation results from direct coupling of radicals by the presence of C oxygenates. Thus in the
[12,59]. The suggestion of Clarke et al. eliminates presence of methanol no change in the ESR signal
the requirement of strongly basic sites for proton was observed when the system transformed from
removal from C–H, a major drawback of a number being nonactive to being active for forming C–C
of introduced mechanisms [12]. However, this bonds. When the methanol feed was stopped a
concept was considered as unlikely because of the clearly observable ESR signal due to organic free
probability of radical interaction with the zeolite radicals appeared, which means that free radicals
surface [11,57], Therefore, Chang et al. suggested are present in the C–C active catalyst, but it is
the following reaction path (Scheme 3). unclear (doubtful ) whether they are essential for
Choukroun et al. [60] observed that radical the reaction or not [64].
dimerization of dimethyl ether gave dimethoxye- Most of the mechanistic studies related to the
thane exclusively (in fluorosulfonic acid ), however, methanol to hydrocarbons reaction were carried
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 11

out using ZSM-5 as catalyst. The mechanisms ethene since this would imply a 12C/13C ratio larger
proposed so far may broadly be classified into than one. The majority of the propene molecules
two groups: were formed directly from methanol [65,66 ].
The consecutive type mechanism with one carbon The nature of the ‘hydrocarbon-pool’ mecha-
from methanol adding during each step. Addition nism was not discussed by Dahl and Kolboe. The
and cracking reactions of the alkene molecules methanol to hydrocarbons reaction could proceed
may take place as illustrated as follows [65]: via a ‘rake mechanism’ as first proposed by
Cormerais et al. [67]. Their mechanism may also
2C C H +H O
1 2 4 2 be consistent with the ‘hydrocarbon-pool’ type
C H +C C H mechanism. In the case of Cormerais et al. [67],
2 4 1 3 6
the hydrocarbon pool would rather consist of
C H +C C H …
3 6 1 4 8 ethoxy, propoxy, butoxy groups etc., formed by
A parallel type mechanism, known as ‘hydrocarbon- methylation of the next lower member. From
pool mechanism’ was suggested by Dahl and results obtained in studies of methanol conversion
Kolboe [65,66 ], who studied the methanol to on H-ZSM-5 [54,68,69] the authors believe rather
hydrocarbons conversion applying SAPO-34 as that the reaction proceeds via an adsorbate which
catalyst and 13C-labelled methanol as feed [and is continually adding and splitting off reactants
(12C) ethene (made in situ from ethanol ) ] (see and products. Their results were in agreement with
Scheme 4). a ‘hydrocarbon-pool’ type mechanism, and ethene
was observed as an inert primary product. In the
co-reaction of ethene and methanol over SAPO-34,
most of the product was made from methanol,
and there was about 90% unconverted ethene
remaining at the time when methanol was fully
converted. The propene and the butenes made
during the reaction exhibit distributions of 12C
and 13C which are nearly random at all times. This
observation supports the ‘hydrocarbon-pool’
mechanism, too. Randomness will, however, also
result if propene and butenes were adsorbed, isoto-
Scheme 4 pically isomerized, and desorbed at a much faster
rate than ethene [65]. Methane was mainly formed
The ‘hydrocarbon-pool’=(CH ) represents an from methanol–dimethyl ether, and only to a
2n
adsorbate which may have many characteristics in minor extent by fragmentation of higher hydro-
common with ordinary coke, and which might carbons. As would be expected, the same applies
easily contain less hydrogen than indicated. It to CO and CO . Some mechanistic details using
2
would perhaps be better represented by (CH ) SAPO-34 may differ from those obtained over
xn
with 0<x<2 [66 ]. H-ZSM-5, due to the unusual high selectivity for
When using SAPO-34 in the title reaction, the the formation of ethene (and propene) [66 ].
product pattern is thus simpler than, e.g. ZSM-5, The distribution of olefins formed in the
where a much wider range of products was found. H-ZSM-5 catalyzed conversions of methanol were
Therefore, it might be easier to obtain a picture of examined at low partial pressures as a function of
the reaction pathway using SAPO-34, and the contact time by Dessau [70]. The results clearly
authors showed that the consecutive mechanism, indicated that ethylene was only obtained at long
in as far as propene formation was concerned, did contact times, and then via secondary reequilibra-
not turn out to be valid. They concluded, that tion of the primary kinetic olefinic products, pro-
only a minor part of the propene molecules may pylene and butenes [70]. Theoretical (NDO)
have been formed by addition of methanol to calculations suggest that the (secondarily) formed
12 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

ethoxyl fragment may be alkylated by methanol Salehirad and Anderson [76 ] reported on the
to a higher fragment or decompose with the use of 13C MAS NMR spectroscopy to monitor
desorption of ethylene [71]. the course of methanol conversion over
The complexity of the methanol to hydro- H-SAPO-34 applying sealed capsules containing
carbons reaction as a result of rapid secondary the catalyst with different loadings of 13C enriched
reactions as well as limitations of suitable tech- methanol. Being isostructural with the mineral
niques to follow this reaction has always been a chabazite, SAPO-34 has a cage structure rather
problem to get conclusive evidence for the chemis- than a channel structure as in ZSM-5. The acidity
try of the methanol conversion. Besides the topic of H-SAPO-34 is lower than that of H-ZSM-5.
of formation of the first C–C bond, the number These two features have consequences with respect
and nature of intermediates have been a matter of to the catalytic performance of SAPO-34 such that
large interest. 13C MAS NMR spectroscopy turned this catalyst is selective for the synthesis of light
out to be a powerful tool with respect to the olefins, i.e. C –C hydrocarbons. The lower acidity
2 4
identification of the intermediates formed, and a and high selectivity of H-SAPO-34 in contrast to
number of investigations has been conducted H-ZSM-5 enable a better understanding of the
mainly applying ZSM-5 and SAPO-34 catalysts. conversion reaction mechanism using 13C MAS
Anderson and Klinowski [72,73] monitored the NMR spectroscopy. The authors stated experi-
catalytic conversion of methanol to hydrocarbons mental evidence for a stepwise methylation reac-
over ZSM-5 at temperatures up to 370°C and tion via surface-bound species derived from: (a)
identified 29 different organic species in the the treatment of several samples with different
adsorbed phase using 13C MAS NMR spectro- loadings of methanol at 250–290°C and; (b) when
scopy. The authors were able to observe the fate either 13C-methanol is coadsorbed with 12C-ethy-
of those species during the course of the reaction. lene over the catalyst or 12C-ethylene is reacted
This technique allowed them to follow different with pre-13C methylated SAPO-34. The hydro-
kinds of shape selectivity in the zeolite host, to carbon products in these experiments were mainly
identify CO as an intermediate in the reaction, and isobutane and isopentane as well as methane,
to distinguish between mobile and attached species. ethylene and propane [76 ]. The presence of
Only methanol was present at room temperature. branched C and C saturated hydrocarbons in
4 5
In samples treated at 150°C dimethyl ether was the intracrystalline space causes an additional
found. At 250°C and above a new signal, due to steric constraint on the diffusion of linear hydro-
CO intermediates, appeared. A number of aliphatic carbons since isobutane and isopentane are too
and aromatic compounds were formed at 300°C large to leave the intracrystalline pore system. This
and above (see Fig. 3 and Table 1) [72,73]. favors the diffusion of C and C species out of
1 2
However, the role of CO in the conversion of the crystallites and is responsible for the apparent
methanol to gasoline on zeolite H-ZSM-5 has been overall selectivity of SAPO-34 for ethylene [77].
debated by Munson et al. [74], based on their in Hydrocarbon formation from methanol on
situ 13C MAS NMR investigations. In contradic- H-SAPO-34 was monitored in a closed static
tion to Anderson and Klinowski, Munson and system by 13C MAS NMR and in an open flow
coworkers concluded that CO is neither an inter- system by gas chromatography by Xu et al. [78].
mediate nor a catalyst in MTG chemistry on Methanol was shown to be converted via the same
zeolite H-ZSM-5 [74]. pathway in either the static or the flow systems
The combination of 13C NMR spectroscopy and but with steady state and non-steady state modes,
gas chromatography allowed Derouane and respectively. The medium acid strength of
coworkers to support the carbenium ion mecha- H-SAPO-34 is thought to be decisive for its good
nism for the conversion of methanol to hydro- selectivity towards the formation of light olefins
carbons over H-ZSM-5 since carbenium ions are [78]. Similar in situ solid-state 13C MAS NMR
very strongly stabilized on the acidic and highly investigations have been performed by other
polarizing surface of this zeolite [75]. authors as well [79–82]. Finally, tritium NMR
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 13

Fig. 3. 13C MAS NMR spectra of a ZSM-5 sample heated to 300°C for 35 min and recorded with proton decoupling only: (a)
aliphatic region; and (b) aromatic and CO region. Intensities in (a) and (b) are not on the same scale. The insert shows J-coupling
of methane and cyclopropane carbons (recorded without decoupling). Spectral assignments are given in Table 1 (reproduced by
permission of the American Chemical Society, Washington DC ).

investigations were carried out in order to monitor carbons reaction. Again a review exists covering
routes to the ethylene production on the conver- the use of spectroscopic techniques to investigate
sion of methanol to hydrocarbons applying the MTG process [91]. Examples are quoted of
H-ZSM-5 as catalyst [83]. catalyst characterization, investigation of the first
Besides solid-state NMR spectroscopy, a variety C–C bond formation, alkene oligomerization and
of other techniques like in situ FTIR spectroscopy catalyst deactivation through coke formation.
[26,84,85], flow reactor/GC-MS spectrometry Obviously, a large number of experiments have
[13,86–88], temperature programmed desorption been done to resolve the question of C–C bond
( TPD) [49,85,89] as well as differential scanning formation from methanol. Although the answer
calorimetry (DSC ) [69,90] have been used to still remains elusive, these experiments tell us at
follow the complexity of the methanol to hydro- least what is probably not involved in the bond
14 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

Table 1
Assignment of 13C MAS NMR signals for the ZSM-5 sample treated at 300°C for 35 min (according to Ref. [73])

Compound Aliphatic Aliphatic Aromatic


signal no.a signal intensityb signal no.a

Isobutane 3 s
Propane 7 s
CO s
n-Butane 3, 9 m
n-Hexane 1, 4, 8 m
Isopentane 1, 2, 4, 10 m
n-Heptane 1, 2, 4, 8 m
Methane 13 m
Ethane 11 w
Cyclopropane 12 w
o-Xylene 5 s 4, 12, 17
1,2,4,5-Tetramethylbenzene 6 s 8, 11
p-Xylene 6 s 7, 13
1,2,4-Trimethylbenzene 5, 6 m 4, 6, 8, 11, 13, 16
1,2,3,5-Tetramethylbenzene 5, 8 m 5, 7, 10, 13
Toluene 5 w 2, 13, 14, 18
m-Xylene 5 w 2, 12, 15, 17
1,3,5-Trimethylbenzene 5 w 3, 16
1,2,3-Trimethylbenzene 5, 8 w 5, 7, 16, 19
1,2,3,4-Tetramethylbenzene 5, 8 w 7, 9, 16

aRefers to Fig. 3.
bs=strong, m=medium, w=weak. Owing to the nuclear Overhauser effect (NOE) these intensities give only a rough guide to the
concentration of each species.

formation, particularly in the presence of zeolite dimethyl ether formation is much faster than the
catalysts [12]. So far, the surface-bound species, subsequent reactions, so that the oxygenates are
that means the carbene and oxonium ylide propos- at equilibrium [11,36 ]. Based on the autocatalytic
als have received significant experimental support, nature of the methanol conversion over ZSM-5,
especially with respect to the methanol conversion Chen and coworkers applied a kinetic model,
over ZSM-5 catalysts. The common feature of which takes into account that the rate of disappear-
both approaches, which is the involvement of the ance of oxygenates is accelerated by their reaction
zeolite conjugate base which acts as a base towards with olefins. At low conversions (<50%) the con-
a non-bonded reactant species, has been ques- version of olefins into aromatics can be neglected
tioned [35]. Especially for the methanol conversion and a reasonable fit was obtained, meaning that
to hydrocarbons over H-ZSM-5, evidence for a the autocatalytic step is much faster than the
mechanism based on the intermediacy of surface- formation of the first olefin from the oxygenates.
bound species (methoxy groups) has been obtained Fitting experimental data obtained on H-ZSM-5
spectroscopically [26,39,84,92–94]. The authors with varying concentrations of acid sites showed
concluded that the surface-bound species is depro- a linear correlation between the rate constant of
tonated via a nearby zeolite basic site (‘adjacent the reaction of oxygenates with olefins and the
Al–O site’) to a surface-bound carbene which in intrinsic acid activity of the catalyst [11,95].
turn engages in C–C bond formation [76 ]. Ono and Mori [51] proposed the first step in
Kinetic investigations related to the methanol to the model of Chen and Reagan [95] to be bimolec-
hydrocarbons conversion normally consider the ular. Assuming this reaction to be of second order
methanol–dimethyl ether mixture as a single with respect to the concentration of the oxygenates
species. This seems to be justified, since the and the autocatalytic reaction to be of first order
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 15

with respect to both the oxygenates and the olefins, A kinetic model for the methanol conversion to
reasonable fits of the kinetic model were obtained olefins with respect to methane formation at low
at 219 and 239°C, indicating that the autocatalytic conversions was discussed in the literature [102].
effect decreases with increasing temperature The study of Hutchings et al. revealed that, at low
[11,51]. conversion, hydrogen formation from methanol
The model of Chen and Reagan [95] was modi- decomposition and the water gas shift reaction
fied by Chang through adding a bimolecular step were significant over the zeolite ZSM-5, and, there-
accounting for the carbene insertion in the primary fore, these reactions should not be included in the
olefins [96 ]. By assuming all reactions to be first derivation of kinetic models for this reaction. In
order and the carbenes to be in steady state, his view if this, the authors considered it unlikely that
model predicts a lower rate constant for the reac- methane formation predominantly proceeds via
tion of the oxygenates with the olefins than for the reaction of carbene intermediates with molecu-
the carbene insertion. This was found to be in line lar hydrogen [102].
with the observation that carbene insertion into a The effect of axial gas dispersion on the light
C–H bond is slower than carbene addition to olefin yield during the MTO reaction was investi-
double bonds. The comparison with experimental gated by Tshabalala and Squires [103] using
data was acceptable. Chang’s description of the pulsed-tracer experiments conducted for a hori-
carbene insertion in olefins predicts a redistribution zontal gas flow through a microreactor. At a given
of the olefinic species, but not a net increase in the WHSV, a large variation in axial Peclet numbers
production of olefins [96 ]. This was corrected for was achieved, with values between 2 and 20 [103].
by Anthony [97] and Ref. [11]. A kinetic model was used by Luk’yanov [104]
A revision of Chang’s modified model [96 ] was as the basis of a computer simulation, establishing
suggested by Sedran et al. [98] who were able to the influence of the correlations among the rates
predict the product distribution at various conver- of the fundamental reaction steps on the composi-
sion levels and temperatures. Anthony and Singh tion of the products formed. Conditions for opti-
[99] concluded from a kinetic analysis of the mized activity of the catalysts for the conversion
methanol conversion to lower olefins on chabazite of methanol were formulated. The influence of the
type zeolites that propylene, methane and propane Si–Al ratio in ZSM-5 zeolites on their activity in
are formed through primary reactions and do not each major reaction step was established [104].
participate in any secondary reactions. However,
dimethyl ether, CO and ethane do. Ethylene and
CO appeared to be produced by secondary reac- 3. Summary of the early studies (before about
2
tions. Furthermore, the product selectivities could 1990)
be correlated to the methanol conversion even
though the selectivity and the conversion changed As already mentioned, a number of reviews exist
with increasing time on stream due to deactivation dealing with the methanol-to-hydrocarbons tech-
by coke formation [11,99]. nology, especially covering the research and devel-
Based on the carbene mechanism, Mihail et al. opment work performed during the 1980s and
[100] developed a kinetic model for the MTO earlier [4–6,11]. The interested reader should con-
reaction up to C H containing 33 reactions. The sult these reviews, however, a brief summary of
5 10
formation of carbene from dimethyl ether turned the earlier work is given in the following.
out to be the rate determining step among the
kinetic parameters evaluated. This is in line with 3.1. Microporous/zeolitic materials
the findings of Chen and Reagan [95] and Ono
and Mori [51]. After all, the model was extended 3.1.1. Large-pore microporous/zeolitic materials
to comprise 53 reactions, including the formation Differently modified Y zeolites have been
of aromatics and C+ aliphatic compounds applied by Salvador and Kladnig [39] in the MTO
5
[11,101]. reaction. Light olefins, mainly propylene, were
16 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

observed at 250°C. The activity of H–Y was high the MTO reaction in the presence of a diluent
compared with Na–Y, which was not active in the selected from water, hydrogen-containing gas,
temperature range studied, i.e. up to 350°C. Ion- nitrogen, He, CO and CO or mixtures of these
2
exchanged Y zeolite samples with less than 16.4 diluents [114].
Na+ and K+ ions per unit cell are active for the Mordenite is a highly active catalyst for the
dimethyl ether conversion, according to Cormerais conversion of methanol to hydrocarbons, prefera-
et al. [105]. Catalytic activity in the methanol-to- bly in the temperature range of 300–500°C
hydrocarbon reaction of NaH–Y zeolite samples [11,115]. However, mordenite deactivates rapidly
was observed at lower temperatures as the degree [116,117]. The selectivity to olefins and the resis-
of proton exchange increased [106 ] and at higher tance to deactivation can be improved by alumi-
temperatures (350°C ) as the dealumination degree nation, however, this treatment reduces the acidic
increased [107]. The temperature dependence indi- character. Propylene is the main olefin produced
cates that increasing proton exchange leads to on dealuminated mordenite. Stable modified mor-
more active sites on Na–Y zeolite, in line with denite catalysts were obtained when the dealumina-
findings by other authors [105]. MTO reactions tion was carried out at about 100°C with 8 M HCl
using dealuminated cubic faujasite (Si–Al ratio of or by Ba2+ exchange and chemical vapor
5.7) were studied at about 400°C by Kubelková deposition of Si from Si(OCH ) [117–119].
34
et al. [108], who detected mainly C –C olefins Ba2+-exchanged dealuminated mordenites are
3 7
besides aromatics up to C . Furthermore, NaH–Y more resistant to deactivation by coke than the
9
(76% ammonium exchange degree) with Si/Al parent dealuminated mordenites. As the degree of
ratios between 2.5 and 8.9 were used as methanol Ba2+-exchange increased, the selectivity to propyl-
conversion catalysts at 300°C, however, mainly ene and butenes was enhanced, while the selectivity
aromatization products were detected under the to aromatics dropped [119]. Some technical
experimental conditions applied [109]. aspects related to the methanol-to-hydrocarbons
A selectivity control to olefins was aimed at by technology are covered by patents using dealumi-
Davidova et al. applying Ni, Cr and Pb exchanged nated mordenite (Si/Al ratio higher than 80) as
Ca–Y zeolites [110]. Ni exchanged Ca–Y produced catalyst [120,121]. Isomorphously substituted erio-
mainly C –C olefins when precalcined in a He nite ( ERI ) and mordenites (MOR) were prepared
2 4
stream. The incorporation of Cr and Pb into and tested in the MTO reaction at 350–450°C
NiCa–Y zeolites changed dramatically the selectiv- using the pulse method in a fluidized catalyst bed
ity to olefins, due to a preferred formation of microreactor. The yield of olefins formed decreased
highly dispersed nickel particles which favor the in the following order:
formation of paraffins. A second group prepared
Y zeolites modified by Pt2+, Pd2+, Zn2+ or Ni2+ B−ERI>ERI>Ga−ERI>Fe−ERI and
cations [111]. A predominant formation of C -C B−MOR>MOR>Ga−MOR>Fe−MOR
2 5
paraffins was found only for catalysts having a
low metal hydrogenation activity. The olefin yields were highest for the boron-
Schwartz and Ciric [112] applied rare-earth and containing zeolites and lowest for the ferrozeolites
Zn-exchanged X-type zeolite in the MTO reaction. [122]. The catalyst lifetime of mordenite in the
In the range of 330–390°C, the MeOH conversion MTG reaction was improved by dealumination,
and the molar selectivity towards C –C olefins on and the activity of highly improved dealuminated
2 4
RE–X zeolite reached 51% and 43.3 mol%, respec- MOR was maintained at least for eight hours. As
tively. Zn–X zeolite was less active. a result, the long-life dealuminated mordenite was
A methanol-rich feed was converted over a obtained by the preferential removal of strong
ZSM-12 zeolite-containing catalyst which con- acidic sites in the internal framework without
tained, in addition, magnesium and/or manganese removing the aluminum atoms on the external
into an olefin-rich hydrocarbon mixture, especially surface [123].
enriched in propylene [113]. ZSM-45 was used in AlPO -5, SAPO-5 and MeAPO-5 molecular
4
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 17

sieves were used in the methanol-to-hydrocarbons related to the required concentration of the inter-
reaction as well [124]. AlPO -5 showed very low mediate dimethyl ether [130]. In a batch reactor,
4
activity compared with SAPO-5. Very high metha- the methanol conversion also showed an induction
nol conversions (almost 100%) were obtained with period, whereas this was not the case with dimethyl
SAPO-5 between 380 and 500°C, however, the ether as feed [131]. The influence of the size and
olefin selectivity decreased with increasing temper- shape of ZSM-5 catalyst particles, the operation
ature, whereas the selectivity towards aromatics conditions and the internal recycle on olefin yield
increased. The highest selectivity to hydrocarbons was investigated for a fixed bed reactor system
was obtained with BeAPO-5 [124]. Tapp et al. [132]. The zeolite was agglomerated by extrusion
[125] applied SAPO-5 and CoAPO-5 in the metha- to cylinders, rings and monoliths. The use of
nol-to-hydrocarbons reaction at 360°C. The typical monoliths as catalysts was not advantageous, due
hydrocarbon fraction contained 45% light olefins, to problems of heat transfer.
15% higher aliphatics and 40% aromatics. IR and The effect of increasing the temperature on the
TPD experiments demonstrate that AlPO s possess product distribution in the methanol-to-hydro-
4
a very low acidity. The introduction of bivalent carbons reaction using H-ZSM-5 is similar to
and tetravalent cations into the framework of increasing the space time at a constant temperature
aluminophosphates results in the generation of [36 ]. At higher temperatures, the olefin formation
Brønsted acid sites, which enhance the catalytic is favored with respect to the formation of aro-
activity for methanol conversion to hydrocarbons matics [128,129]. Thermodynamically, the effect
[125–127]. of increasing the temperature over the same range
of partial pressures of olefins is to shift the distribu-
3.1.2. Medium-pore microporous/zeolitic materials tion towards the lower olefins [24]. This equilib-
The pentasil-type zeolite ZSM-5 has been the rium distribution is only approached at low
candidate among the medium-pore zeolites to be conversions of methanol to hydrocarbons [11].
investigated as catalyst for the methanol-to-hydro- Whereas ethylene is most abundant at low temper-
carbons reaction. A very large number of papers atures, butene and especially propylene become
exist dealing with this microporous material as more important at higher temperatures [133]. Ono
MTO or MTG catalyst. A systematic literature et al. [134] monitored an abrupt jump in the
survey has been prepared by Froment et al. [11], conversion of methanol to hydrocarbons applying
especially with respect to this zeolite type, where H-ZSM-5 at constant space time when increasing
certain aspects related to the catalytic test condi- the temperature from 280 to 300°C. It was con-
tions (space time, temperature, partial pressure, cluded, that the reason for this jump was the onset
etc.) as well as to the catalyst preparation (ion of cracking of oligomers which were formed from
exchange method, isomorphous substitution, Si/Al small olefins and remain in the ZSM-5 pores at
ratio of the material, etc.) have been investigated. low temperatures [11,130]. Jansen van Rensburg
The effect of space time on the methanol conver- et al. [135] concluded that at low conversion
sion and the hydrocarbon distribution clearly and/or low temperature, ethylene is mainly a resid-
demonstrates that C -C olefins are intermediates ual primary reaction product. At higher temper-
2 4
in the MTG reaction [36,128,129]. The yield of atures, however, ethylene is a product resulting
light olefins increases with space time and reaches from secondary cracking [11].
a maximum, indicating the intermediate character The partial pressure of methanol or dimethyl
of the light olefins formed. An important feature ether has been shown to have a strong influence
of the methanol to hydrocarbons reaction on on the olefin selectivity. A decrease of the partial
ZSM-5 is that a further increase of space time pressure tends to enhance the olefin formation and
results in a continuing change of the hydrocarbon suppress the aromatization reactions [136 ]. Chang
distribution, although the conversion is complete and Lang [137] observed an eight-fold increase in
[11]. The sudden increase in the conversion of selectivity towards light olefins when the partial
methanol to hydrocarbons with space time was pressure of methanol was changed from 1 to
18 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

0.07 bar and keeping the space time constant. All by Suzuki et al. [149]. The average diameter of
the results of the experiments at low partial pres- the zeolite crystallites increased as the water
sures of methanol indicate that ethylene is not the content in the reaction mixture increased. The
initial olefin formed from methanol, but is formed authors concluded that lowering the aluminum
by the secondary re-equilibration of the primary content at the outer shell of the crystallites by
olefins, i.e. propylene and butenes. The ethylene lowering the water content of the reaction mixture
formation at higher partial pressures of methanol depressed the coke deposition and improved the
is mainly the result of a thermodynamic control catalyst stability [149]. Other modifications of
[11]. This is in fact in contradiction to Haag et al. ZSM-5 zeolites ( like ion exchange [150,151] and
[138] who applied a high-silica H-ZSM-5 catalyst others) are reported as well:
(Si/Al ratio of 800) at atmospheric pressure and Following Inui et al. [152], acidification of
low methanol conversion and concluded that ethy- ZSM-5 should be done by ammonium chloride or
lene is the primary olefin from methanol. Increased HCl in order to obtain higher olefin selectivities,
selectivities towards the formation of ethylene by because less strongly acid sites are formed in this
decreasing the methanol partial pressure were way than by ion exchange with ammonium nitrate.
observed by several authors [139–141]. Similar investigations were performed by other
Investigations related to the conversion of meth- groups, partly arriving at the same conclusions
anol as a function of space time on H-ZSM-5 [153–155].
catalysts with varying Si/Al ratios showed that the Several modifications of the Mobil type ZSM-5
maximum in the yield of the intermediate light zeolite were suggested in order to improve the
olefins shifted towards higher methanol conver- selectivity towards light olefins in the MTO reac-
sions with increasing Si/Al ratio [128,133]. Wu tion, especially with respect to ion exchange and
and Kaeding [142] correlated the product distribu- impregnation methods: Cs+ and Ba2+ exchange
tion of the methanol to hydrocarbons reaction on [133,147,150,151], phosphorus impregnation
ZSM-5 with the conversion of the alcohol. [133,156–161], Mg modification [146,160,161],
Ethylene was the major primary hydrocarbon pro- Mo modification [159], Sb O impregnation [162]
2 3
duced from methanol at low conversion. At higher and adsorption of Si-containing compounds
conversion, the olefins undergo scrambling or [163,164].
thermodynamic equilibration reactions also pro- Isomorphous substitution of Si by Al or other
ducing olefin mixtures with good correlation metals results usually in the generation of catalyti-
coefficients for the Flory equation. cally active sites. Especially B-ZSM-5 type catalysts
Several authors report on the influence of the have attracted a lot of attention with respect to
synthesis conditions of H-ZSM-5 on the selectivity the conversion of methanol [165–174]. It was
towards light olefins, including the effect of concluded that B-ZSM-5 is more suitable for the
sodium-free gels (higher selectivity towards light production of propylene, while ethylene is more
olefins) and the crystallization time (decrease in selectively formed on Al-ZSM-5 as well as on
selectivity at longer crystallization times) [143– fluorinated aluminosilicate pentasiles [11,175].
146 ]. The crystal size of ZSM-5 zeolites has been ZSM-5 catalysts containing Zr, Hf and Ti are also
found to be an important factor influencing the more selective in the conversion of methanol to
methanol conversion to light olefins. Applying olefins [176–178]. Inui et al. [179] introduced a
crystals with a size of less than 2 mm, a higher procedure for preparing various high-silica ZSM-5
selectivity for the production of light olefins and zeolites by replacing Al at the stage of gel forma-
particularly ethylene has been reported tion by transition metals. Ti and Zr containing
[139,147,148]. zeolites were the most selective ones towards light
The effects of the water content of the reaction olefins. V-ZSM-5 with a high vanadium content
mixture on the physicochemical properties and the (Si/V ratio of 90) revealed a high selectivity
performance of ZSM-5 type zeolite catalysts in towards light olefins [180].
methanol conversion reactions were investigated Isomorphously substituted ZSM-5 catalysts
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 19

were investigated in detail by Inui et al. [181,182] Table 2


in order to optimize the selectivity of metal con- Patent literature covering the improvement of the ZSM-5 cata-
lyst and/or process conditions in the methanol-to-hydrocarbon
taining zeolites in the MTO reaction towards ligh- reaction (time period before 1989)
ter olefins. At complete conversion of methanol,
the selectivity towards olefins on the various metal Improvement/process attempt Refs.
substituted ZSM-5 zeolites varied as followed:
Fluidized bed reactor technology [186–189]
Interstage sorption fractionator technology [190–192]
Ga=Cr<V<Ge<Mn<La=Al<Ni<Zr
Adiabatic reactor technology [193]
=Ti<Fe<Co=Pt Different reaction zones with recycling [194]
Presence of diluent [195,196 ]
It was clear that ZSM-5 substituted with Fe, Co Aromatics-rich hydrocarbon mixtures [197–200]
Large catalyst particles [201]
and Pt revealed the highest selectivity in the forma- Reactivated zeolite [202]
tion of light olefins. The catalytic properties of Zeolite catalyst in a silica binder [203]
isomorphously substituted Be-ZSM-5 were studied Zeolite catalyst modified by metal inclusion [204]
for the methanol to hydrocarbons reaction by Ga-ZSM-5 [205]
Pentasil LZ 40 [206,207]
Romannikov et al. [183], concluding with the
Temperature and pH variations [208]
preferred formation of olefins for this type of
catalyst compared with Al-ZSM-5, which pro-
duced mainly paraffins and aromatics [183].
However, the methanol conversion was consider- dences obtained can be used to predict the hydro-
ably higher with Al-ZSM-5 than on the Be-ZSM-5, carbon composition of gasoline with account taken
except for one example. for monitoring and control of the necessary degree
Fleckenstein et al. [184] compared the conver- of acidity of the high-silica zeolites both at the
sion of methanol to olefins between H-ZSM-5 and stage of preparation of the catalysts and during
H-ERI, concluding with a preferred formation of their use [209].
lower olefins using H-ERI, whereas H-ZSM-5 pro- Besides ZSM-5, other medium-pore type zeolites
duced larger amounts of aromatics. Faster deacti- have been investigated with respect to the catalytic
vation was observed for the small-pore erionite conversion of methanol to hydrocarbons, however,
zeolite [185]. much less attention has been paid to the other
A number of patents exist and has been reviewed end-member of the pentasil series, i.e. ZSM-11,
with respect to the improvement of the ZSM-5 probably because it was soon recognized that the
catalyst and/or the process conditions in the meth- shape-selective properties of ZSM-5 were superior
anol to hydrocarbon reaction since this zeolite has to those of ZSM-11 [11,210,211]. In the conver-
attracted a lot of interest during the last decades. sion of methanol to gasoline on ZSM-11, much
The available information is quite tremendous, less C –C products were observed than with
1 2
and the results of this search have been summa- ZSM-5 while C+-aliphatics are more abundant.
6
rized in Table 2 (covering the time before 1989). Whereas on ZSM-5 the aromatics fraction is
However, details should be followed up in the mainly composed of xylenes, on ZSM-11 more
original patent literature. C –C aromatics are produced [11]. The differ-
9 10
An examination of published experimental data ence in shape-selectivity is attributed to the relative
for a wide range of concentrations of strong acid dimensions of the channel intersections and to the
centers of zeolite ZSM-5 made it possible to estab- difference in the relative length of the straight
lish the ultimate nature of dependences of the yield ( ZSM-5, ZSM-11) and tortuous channels
of individual products of methanol transformation ( ZSM-5). Furthermore, since ZSM-11 contains
into gasoline [209]. Regression equations describ- only one type of channel, the molecular traffic
ing the relationships between the concentration of control shape-selectivity should, in principle, not
acid centers, the Si/Al ratio, and the composition occur in this catalyst. This provides a possible
of synthesis products were formulated. The depen- explanation for the higher yield of polysubstituted
20 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

aromatics when using ZSM-11: the methanol reac- Hydrocarbon fractions with more than 80 wt% of
tant and aromatic products cannot avoid counter- light olefins were obtained with a dealuminated
diffusion as in ZSM-5, and this increases the H-erionite [4,11,215,216 ]. Dealuminated chaba-
probability of alkylation of the aromatics [11]. zite-type catalysts (Si/Al ratio of about 10) gave
Ione et al. [165] studied the properties of molar selectivities towards light olefins of up to
ZSM-11 type metallosilicates in the conversion of 94% at 100% methanol conversion, according to a
methanol to olefins: Al-ZSM-11 proved to be the study performed by Cartlidge and Patel [217].
most active and selective catalyst in this reaction Wunder and Leupold [218] applied a mixture of
compared to other metal-substituted ZSM-11 sys- Mn-exchanged chabazite and erionite zeolites,
tems. Chang et al. [212] patented the application obtaining a hydrocarbon fraction containing
of ZSM-48 (high silica medium-pore zeolite) for 37.1% ethylene, 26.5% propylene and 2.7% butene
the production of olefins from methanol or at 90% conversion of methanol (feedstock —
dimethyl ether. At high conversion and at temper- 70:30 vol% mixture of methanol–water; reaction
atures lower than those needed to obtain compara- temperature, 400°C ). The selectivity towards ethy-
ble product distributions with ZSM-5, the reactor lene and propylene was higher than when normal
effluent consisted mainly of olefins. The higher H-forms of these zeolites were applied. The same
olefin selectivity obtained with ZSM-48 are proba- group compared the chabazite type zeolites with
bly taken as conditions optimal for olefin pro- pentasils, concluding with a smaller tendency of
duction on ZSM-48 and for gasoline production the latter zeolites to form coke, whereas the small-
on ZSM-5. Yet, from the product distribution it pore zeolites revealed a higher olefin selectivity
can be deduced that ZSM-5 is far more active than [219]. Cation exchanged chabazites were used by
ZSM-48. The major difference is found in the Singh and Anthony [220], who obtained yields of
production of higher molecular weight olefins, ethylene, propylene and propane in the range of
which is significantly higher on ZSM-48. In addi- about 35, 30 and 25%, respectively, in the metha-
tion, fewer aromatics are formed on ZSM-48. The nol-to-hydrocarbons reaction. The deactivation by
high yield of paraffins in conjunction with the low coke was fast [221].
yield of aromatics is indicative of severe coke Kljueva et al. [122] evaluated the acidic and
laydown. Indeed, the deactivation of ZSM-48 with catalytic properties of isomorphously substituted
its one-dimensional pore structure was much faster erionite implementing B, Ga and Fe atoms. The
than that of ZSM-5 [11,212]. creation of new acid centers led to catalysts with
Zeolite EU-2, patented by Casci et al. [213], is a lower hydrogen transfer capacity, resulting in
another medium-pore high silica zeolite. When enhanced selectivities towards lighter olefins due
synthesized with a Si/Al ratio of at least 120, to a lower olefin conversion into paraffins. TMA
ZSM-48 is isostructural with zeolite EU-2 [214]. and H-offretite calcined at 450°C have been
A high selectivity towards propylene and butene reported with a certain activity to mainly form
and a low selectivity towards aromatics were propylene, propane, n-butane and n-butene in the
reported in the conversion of methanol to hydro- conversion of methanol into hydrocarbons [222].
carbons on zeolite EU-2 [11,213]. The methanol conversion on zeolite H–T
(intergrowth of erionite–offretite) was monitored
3.1.3. Small-pore microporous/zeolitic materials by Ceckiewicz at 400°C: the products were mainly
Chang et al. [215] claimed in a patent the ethylene, propylene, propane, C -C hydrocarbons
4 5
discovery of chabazite, erionite, zeolite T and and dimethyl ether with a methanol conversion of
zeolite ZK-5 as active catalysts for the conversion 84.5% [85,223,224]. Corresponding results were
of methanol to light olefins. The light olefin con- obtained by Gubish and Bandermann [225], using
centration in the hydrocarbon fraction was always the same type of zeolite.
less than 60 wt% at 100% methanol conversion. A number of patents exist dealing with the
The hydrocarbon fraction became richer in light manufacture of light olefins from methanol con-
olefins as the conversion of methanol decreased. taining feedstocks, applying erionite, TMA-offret-
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 21

ite, zeolite T, clinoptilolite or ZSM-34 as catalysts further characterized by improved life time before
[226–231]. Givens et al. [226,227] claim that the deactivation through coke formation. Deactivated
use of steam as diluent enhanced the selectivity for catalysts are fully restored through air regenera-
ethylene. Hydrocarbon fractions containing up to tion. The crystallinity of the molecular sieve is
90 wt% of light olefins were reported for ZSM-34. retained during methanol conversion and regenera-
Inui and Takegami [232] and Inui et al. tion [239,248]. SAPO-16 and SAPO-44 have been
[233,234] have studied the influence of the crystal- used as catalysts for the conversion of methanol
lization conditions of ZSM-34 as catalyst for the to hydrocarbons as well [240,243,249]. However,
MTO reaction. They found that an optimum selec- the methanol conversion was only 51% at 425°C
tivity was obtained with ZSM-34 crystallized for in the case of SAPO-16. Furthermore, Kaiser
25–39 days at 100°C. applied metalloaluminophosphates (MeAPO) and
Modified clinoptilolite zeolites were applied by metallosilicoaluminophosphates (MeAPSO) as
several authors, both ion-exchanged and natural well in order to form light olefins from a 70:30 wt%
samples, which showed lower activity than the water–methanol mixture [250,251]. MgAPO-34,
former ones [235,236 ]. Hydrocarbon fractions CoAPSO-34 and MnAPO-34 can completely con-
containing up to 84.6 wt% of lower olefins were vert methanol, but methane and carbon dioxide
observed at almost 100% methanol conversion. selectivities were high, indicating that sintering
FU-1 zeolite was used both in its sodium and occurred during pretreatment of the samples [11].
H-forms. Moderate light olefin selectivities were CoAPO-34 showed the following product distribu-
obtained for H-FU-1 zeolite at 450°C: 10 wt% tion (in mol%) at 425°C [ WHSV (methanol ):
ethylene, 26 wt% propylene and 38 wt% butenes, 0.94 h−1, reaction time: 4.8 h]: ethylene 45.3%,
besides paraffins. However, the sodium form of ethane 0.8%, propylene 27.1%, butenes 8.3%,
FU-1 showed negligible activity in the methanol- methane 6.2%, CO 10.0%, traces of propane and
2
to-hydrocarbons reaction [237]. C –C hydrocarbons (olefin selectivity: 80.7%)
5 6
Sigma-1 possesses a high activity for methanol [251].
conversion and a high selectivity for light olefins
at 400°C: the main products obtained were ethy- 3.2. Non-microporous/non-zeolitic materials
lene and propylene. The hydrocarbon fraction
consisted of up to 78% of lower olefins (19% Most of the work concerning the conversion of
ethylene, 46% propylene and 13% butenes) whereas methanol to hydrocarbons has been done on
no aromatic compounds were detected in the pro- microporous/zeolitic materials. However, the con-
duct [238]. version of dimethyl ether and methanol was inves-
Highly selective conversion of light olefins has tigated over several salts of 12-tungstophosphoric
been reported by Kaiser [239–243] using SAPO acid, H PW O , at 290°C [252]. For acid salts
3 12 40
molecular sieves. Especially SAPO-17 and of Na, the catalytic activity decreased in parallel
SAPO-34 have shown a very high selectivity for with the acidity as the Na content increased.
light olefins at almost 100% methanol conversion However, the carbon number distribution and the
[239,243–246 ]. The high activity is related to the olefin/alkane ratio of the resulting hydrocarbon
Al-OH group with a vibration frequency at about mixture were almost independent of the Na
3600 cm−1 [247]. With SAPO-34 combined molar content. These results indicate that the acidity
selectivities to light olefins up to about 96% at determines the activity, but its effect on the selectiv-
complete methanol conversion to hydrocarbons ity is not significant. Light olefin selectivities in the
were observed [temperature 375°C–450°C, atmo- range of up to 65% have been reported by the
spheric pressure, WHSV (methanol ) 0.83– authors [252].
0.87 h−1]. Very low yields of methane and other Moffat [253] studied the methanol conversion
saturated hydrocarbons were found. Increasing the to hydrocarbons on 12-tungstophosphoric acid
temperature results in a smooth increase in selectiv- compounds (H PW O ) as well. The selectivity
3 12 40
ity to ethylene (as high as 61%). The catalysts are to hydrocarbons other than methane was found
22 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

to depend on the reaction temperature and the tions: temperature 300°C, LHSV 0.2 h−1, feed:
residence time as well as on the pretreatment methanol–helium containing 26.2 mol% methanol,
conditions. It turned out that at low residence methanol conversion of 77%) [256 ].
times the predominant product was dimethyl ether
while at high residence times hydrocarbons larger
than methane were evident [253]. The conversion 4. Recent studies of the methanol-to-hydrocarbons
of methanol to hydrocarbons on Aerosil sup- reaction
ported Ag (SiW O ) catalysts were followed by
4 12 40
Ehwald et al. [254] who showed that the support The work published in the present decade is
had a stabilizing effect on the catalytic activity. reviewed in this section, which covers the more
Best results with respect to the production of recent developments related to the topic of this
olefins and other hydrocarbons were obtained in review. However, the ‘border line’ between the
the medium range of surface concentrations of the literature covered in Section 3 (Summary of the
active component. Their results indicated that the early studies) and the present section is not strictly
deposition of coke causing catalytic deactivation a certain year rather than the content of suitable
took place predominantly at the exterior surface publications and their relations to each other.
of the catalyst, whereas active sites within the pore Nevertheless, a rough distribution in earlier and
system remained untouched by deactivation [254]. recent studies should help the reader to follow the
The same group reported on the investigation main lines of the development. The same splitting
of methanol to hydrocarbons using alumina – has been chosen as in Section 3.
supported H (SiW O ) catalysts [255]. Two
4 12 40
different types of (SiW O )4− anions exist on the 4.1. Microporous/zeolitic materials
12 40
surface of the catalysts, depending on the concen-
tration of the heteropoly compound. Increasing 4.1.1. Large-pore microporous/zeolitic materials
the H (SiW O ) concentration from 5 to 40 wt% Following the recent literature there is no doubt
4 12 40
of W results mainly in an increase in the surface that the focus concerning investigations of cata-
density of Brønsted acid sites, whereas the distribu- lysts suitable for the methanol conversion to
tion of their acid strength is only slightly shifted hydrocarbons has been on the small- and medium-
towards higher values. From the dependence of pore-type microporous/zeolitic materials. How-
the yield of the hydrocarbon on the H ever, a few studies have been published dealing
4
(SiW O ) content of the catalysts, it was con- with large-pore type zeolites or other microporous
12 40
cluded that under the conditions applied hydro- materials.
carbon formation from methanol occurred only A process for converting a feedstock consisting
on adjacent acid groups, the density of which was of a mixture of methanol–water (molar ratio of
evaluated to be approximately 1 mmol m−2 [255]. 1:4) to light olefins using ECR-1 (complex syn-
However, in both publications the formation of thetic zeolite having structural similarities to both
ethylene and propylene under these conditions naturally occurring mordenite and mazzite),
(reaction temperature 300°C, LHSV 0.1 h−1, feeds- mazmorite (consisting of complex intergrown syn-
tock: methanol–nitrogen in a molar ratio of 1:1, thetic permutations of ECR-1) and/or ECR-18
methanol conversion higher than 76%) was only (synthetic paulingite-type zeolite) has been pat-
minor compared with the production of C+ com- ented by Janssen and Vaughan [257]. However, at
4
pounds including at large part of aromatics 450°C and a WHSV of 1.0 h−1, the obtained olefin
[254,255]. selectivities were in the range of 27 wt% for ethy-
H PW O catalysts have been patented both lene and 11 wt% for propylene ( ECR-1) as well as
3 12 40
as neat as well as lanthanide metal modified sys- 38 wt% for ethylene and 30 wt% for propylene
tems, and typical olefin selectivities for ethylene, ( ECR-18).
propylene and butenes were 18.1, 21.9 and 11.0%, Higher selectivities to light olefins were obtained
respectively (catalyst: H PW O , reaction condi- by applying dealuminated mordenite (Si/Al ratio
3 12 40
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 23

of 80) in the conversion of methanol to hydro- SAPO-44 and MeAPSO-44 (Me=Co, Mn, Cr)
carbons at 480°C using a water–methanol mixture samples are about three times higher than the
(feed ratio – 70:30 wt%): up to 81% with a metha- selectivities of SAPO-34 and MeAPSO-34 (Me=
nol conversion of 97% [258]. The durable activity Co, Mn, Cr) [259].
of the dealuminated mordenites and the selectivity
to light alkenes can also be improved by decreasing 4.1.2. Medium-pore microporous/zeolitic materials
the partial pressure of methanol and adding more The methanol to gasoline conversion over
water to the feed. The decrease of the partial H-ZSM-5 has been studied by using the ‘temper-
pressure of methanol leads to a higher yield of ature programmed surface reaction ( TPSR)’ tech-
light alkenes and a lower yield of alkanes, oligo- nique [260,261]. This technique is able to delineate
mers and aromatics. Water competes with light the two steps in the process: the dehydration of
olefins for the Brønsted and Lewis acid sites. The methanol to dimethyl ether and the subsequent
adsorption of water on these acid centers reduces conversion of dimethyl ether to hydrocarbons. The
their strength and concentration and, thereby, the activation barriers associated with each step were
probability of their interaction with hydrocarbons. evaluated from the TPSR profiles and are 108 and
As a consequence, the initial conversion of light 195 kJ mol−1, respectively. The methanol desorp-
olefins, mainly propylene, into oligomers, aro- tion profiles showed considerable changes with the
matics and coke decreases. Both the lower yield of amount of methanol molecules adsorbed per
coke precursors and the weaker interaction Brønsted site of zeolite (see Fig. 4 which shows a
between coke precursors and acid sites reduces the typical TPSR of three methanol molecules
deactivation by coke. At the early stages of the adsorbed per Brønsted site of H-ZSM-5).
reaction, the gradual reduction of the number of Furthermore, the authors compared this metha-
acid sites decreases the probability of their inter- nol to hydrocarbons reaction (over H-ZSM-5)
action with light alkene molecules. Thereby, the with the corresponding gas phase reaction, and
conversion of light alkenes into alkanes and aro- the results are shown in Fig. 5.
matics decreases and the selectivity to light alkenes Fig. 5 shows that, once the protonated cluster is
increases [258]. formed, the activation energies for the formation
The acid strength and catalytic activity of of dimethyl ether as calculated for the gas phase
SAPO-44 and the isomorphously substituted and as measured for H-ZSM-5 are comparable in
MeAPSO-44 (Me=Co, Mn, Cr, Zn, Mg) molecu- magnitude, viz. 111 and 107.5 kJ mol−1, respec-
lar sieves in the methanol to hydrocarbons reaction tively. The main difference between the gas phase
were measured by Hočevar et al. [259]. In addition, and the catalytic processes lies in the energetics of
the results were compared with those of AlPO -5 the formation of the protonated methanol cluster:
4
and AlPO -14 samples. The acidity strength of the while the heat of formation of the active catalytic
4
samples has been determined qualitatively from species [(CH OH ) H+-ZSM-5−] is exothermic, the
3 n
the temperatures of dimethyl ether desorption gas phase heat of formation of (CH OH ) H+ is
3 n
peaks, and the following sequence has been strongly endothermic [260,261].
obtained: Density functional theory calculations have been
used to determine the transition states and adsorp-
MnAPSO−44>CoAPSO−44~ZnAPSO−44 tion complexes of reactants, intermediates and
>MgAPSO−44>CrAPSO−44 products as well as the corresponding activation
barriers and adsorption energies of the numerous
>SAPO−44>AlPO −5 steps involved in the methanol to hydrocarbons
4
~AlPO −14 reaction on an acidic zeolite [262]. Two different
4 mechanisms with respect to the formation of
The highest selectivity to ethylene formation dimethyl ether from methanol have been consid-
among these catalysts has been found for ered: the first proceeds via an associative inter-
CoAPSO-44. The selectivities to ethylene of action between two methanol molecules,
24 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

Fig. 4. Temperature programmed surface reaction (TPSR) of methanol over H-ZSM-5. The profiles are for methanol (m/e 31) (Ω),
dimethylether (m/e 45) (–––), water (m/e 18) (#), C -C aliphatic hydrocarbons (m/e 43) ($) and aromatic hydrocarbons (m/e 91)
3 5
(%). The solid line is the fitted profile for water evolution (reproduced by permission of Chemie, Weinheim).

generating directly dimethyl ether, while the other Upon increasing the alcohol pressure up to 1 mbar,
proceeds via a methoxy surface species intermedi- protonized clusters of about three strongly
ate. The presence of water lowers the activation hydrogen bonded molecules were observed. During
barrier of the last mechanism by more than temperature desorption most of the protonated
50 kJ mol−1 [262]. Furthermore, the effect of methanol desorbed without reaction below 250°C.
making the zeolitic cluster slightly more acidic (by Subsequently, two types of methoxy groups were
lengthening the Si–H bond distances) over the present, which were generated by reaction of meth-
activation barriers of dimethyl ether formation has anol with Si-OH-Al and with terminal silanol
been studied using density functional theory calcu- groups, respectively. Above 200°C, dimethyl ether
lations. Changes in the order of 5 kJ mol−1 were was produced from reactions of desorbing metha-
observed [263]. nol with methoxonium ions or methoxy groups
The adsorption structures and thermally induced [264].
surface chemistry of methanol adsorbed on Chinese researchers suggested that lower olefins
H-ZSM-5 was studied by means of IR spectro- mainly originated from the cracking of higher
scopy, mass spectrometry and thermogravimetry olefins (or their precursors) formed through an
conducted by Mirth and Lercher [264]. At low autocatalytic homologation reaction when apply-
equilibrium pressures (up to 10−3 mbar) methanol ing H-ZSM-5 in the MTO reaction [265], whereas
interacts selectively with the Brønsted acid sites of a ‘new zeolite’ developed at the Nanjing University
the zeolite by forming a methoxonium ion which showed high activity, good selectivity and stability
can be oriented in two different forms. The meth- in the MTG reaction [266 ]. However, the structure
oxonium ion points either with its methyl or with of the ‘new zeolite’ was not disclosed by the
its OH+ group to the SiO-Al site of the zeolite. authors!
2
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 25

these results indicate that deactivation of the zeo-


lites is attributable to the poisoning of the outer
surface by coke deposition. From the fact that the
rate of coke formation is independent of the zeolite
crystal size, it was concluded that coke or its
precursors are formed in the zeolite crystal and
migrate to the outer surface of the zeolite crystal.
Finally, the acidic properties (investigated by
means of ammonia-TPD and diffuse reflectance
FT IR spectroscopy) of the calcium containing
aluminoborosilicates synthesized in the presence
of boric acid were similar to those of calcium
containing aluminosilicates synthesized in the
absence of boric acid [267,269].
ZSM-5 type zeolites containing alkaline earth
metals were hydrothermally prepared from synthe-
sis mixtures containing alkaline earth metal salts.
By this method, the amount of alkaline earth
metals occluded in ZSM-5 was more than that of
the alkaline earth introduced by the conventional
Fig. 5. One-dimensional potential energy diagram for the pri-
mary step of the MTG process (methanol to dimethylether and
ion exchange method [270]. The chemical state of
water). The top part of the figure is for the gas phase reaction, the alkaline earth metals in these systems were
whereas the bottom half is the reaction over H-ZSM-5. All studied by diffuse reflectance FT IR spectroscopy,
figures are in kJ mol−1 (reproduced by permission of Chemie, and the spectra of the OH regions are shown
Weinheim). in Fig. 6.
The IR spectrum of H-ZSM-5 was characterized
The catalyst life and performance of calcium by two well-defined signals at 3605 and
containing aluminoborosilicates with various crys- 3740 cm−1, assigned to the acidic bridged OH of
tal sizes for methanol conversion to light olefins Si(OH )Al and to the terminal SiOH, respectively.
were studied by Kawamura et al. [267–269]. The The broad peak at near 3500 cm−1 is attributed to
crystal size was controlled by adding boric acid. hydrogen bonding between adjacent hydroxyl
For the zeolites with similar Si/Al and Ca/Al ratios, groups. Except for Mg-ZSM-5, intensities of all of
the initial catalyst activity was independent of the these signals for alkaline earth metal containing
crystal size and was the same as that of the calcium ZSM-5 were very weak as compared with those of
containing aluminosilicates synthesized in the H-ZSM-5. It was found that these alkaline earth
absence of boric acid which exhibited the best metal containing ZSM-5 zeolites revealed a high
catalyst performance for production of light olefins selectivity to light olefins at temperatures above
from methanol. However, the catalyst life of the 500°C, where ethylene and propylene were formed
calcium containing aluminoborosilicates, defined in yields greater than 50%. The catalyst life of
as the time during which methanol conversion is these systems could be greatly extended by modifi-
100% with no residual dimethyl ether and with cation with alkaline earth carbonates to suppress
light olefin yields of >50%, was extended by a coking and dealumination. Furthermore, it was
decrease of the crystal size. The amount of coke found that calcium phosphate modified micro-
deposited on the zeolite after the reaction increased crystalline H-ZSM-5 synthesized from the synthe-
with decreasing crystal sizes. The rate of coke sis mixtures with low H O/SiO ratios showed an
2 2
formation was independent of the crystal size. As excellent catalyst performance [270].
the decrease in the crystal size corresponds to the Howden et al. investigated a number of factors
increase in the outer surface per gram of zeolite, influencing the light olefin selectivity during the
26 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

Table 3
Methanol-to-gasoline hydrocarbon yields for different reactor
designs (according to Ref. [272])

Fixed bed Fluidized bed


reactor reactor

Hydrocarbon product [wt %]


Light gas 1.3 4.3
Propane 4.6 4.4
Propylene 0.2 4.3
i-Butane 8.8 11.0
n-Butane 2.7 2.0
Butenes 1.1 5.8
C+ Gasoline 81.3 68.2
5
C+ Gasoline 83.9 91.2
5
(including alkylate)

Gasoline octane numbers


RON 93 95
MON 83 85

in a cost advantage over fixed bed MTG of at


Fig. 6. FT IR Spectra of various zeolites (SiO /Al O =200)
2 2 3
(reproduced by permission of The Japan Petroleum Institute,
least 10% [272].
Tokyo). Catalytic conversion of methanol to light olefins
was monitored using high silica zeolite of the
pentasil type MFI structure by Al-Jarallah et al.
conversion of methanol over ZSM-5 [271]. The [273]. The reaction was carried out in a fixed bed
authors found that the most important factors reaction set-up at 400°C, WHSV=4 h−1, pressure
influencing olefin selectivity are the temperature of of 1 bar and a methanol–nitrogen weight ratio of
the reaction and the Si/Al ratio of the ZSM-5. For 2.78. The silicalite zeolite was modified by impreg-
small laboratory scale experiments, they found it nation with metal nitrates of Ag, Ca, Cd, Cu, Ga,
feasible to achieve a light olefin selectivity around In, La and Sr to study their effects on the activity
65%, of which about 15% would be ethylene, by and selectivity of the catalysts to light olefins.
adopting the following parameters (approximate Incorporation of La and Ag led to an improvement
values): in light alkene selectivity of the silicalite by 18%
Partial pressure of methanol 0.2 bar and 14%, respectively (see Fig. 7). This was attrib-
Reaction temperature 450°C uted to enhanced shape selectivity of the silicalite
SiO /Al O mol ratio of ZSM-5 250:1 resulting from reduction of the apparent pore size
2 2 3
Crystallite size <5 mm of the zeolite channels. The active life-on-stream
Mobil published a comparison of their fixed bed was slightly decreased due to coke deposition
versus fluid bed operation results based on the [273].
MTG demonstration plant (4000 t year−1) using Again a number of recent patents exist covering
ZSM-5 as catalyst. The MTG hydrocarbon yields the improvement of the ZSM-5 catalyst and/or
are summarized in Table 3 [6,272]. process conditions in the methanol-to-hydro-
As shown in Table 3, fluid bed MTG gives carbons reaction. The results of this search have
higher yields of higher quality gasoline than fixed been briefly summarized in Table 4, however,
bed MTG. This yield and quality advantage cou- details can be found in the original patent
pled with the improved energy utilization results literature.
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 27

Fig. 7. Product distribution for various modifications of silicalite (ZSM-5) at 400°C, WHSV 4 h−1 and 2.78 (wt/wt) methanol-to-
nitrogen ratio (reproduced by permission of Elsevier Science B.V., Amsterdam).

Finally, H-Theta-1 zeolite ( TON ) has been pat- medium-pore zeolites was tried through modera-
ented as a catalyst for the methanol-to-hydro- tion of their strong acid sites by incorporation of
carbons reaction, converting a methanol/1-butene basic materials like Ca, Mg, Zn, Sr, B and P. As
mixture (82:18 mol% ratio) at 400°C preferentially the selectivity to olefins increased but the activity
to isopentenes [281,282]. decreased, the reaction temperature was raised up
In conclusion, an increase in olefin selectivity of to 600°C from around 400°C. Unfortunately, on
these catalysts and at those higher temperatures,
Table 4 the major olefin in the products was propylene,
Recent patent literature covering the improvement of the and the most valuable ethylene was formed fairly
ZSM-5 catalyst and/or process conditions in the methanol-to- little.
hydrocarbon reaction (1989 or later)

Improvement/process attempt Refs. 4.1.3. Small-pore microporous/zeolitic materials


Since the beginning of the 1990s, the main focus
Fe, Al containing ZSM-5 [274]
concerning the catalytic system to be investigated
Formulated pentasil type zeolite [275]
Olefinic gasoline and distillate range liquids [276 ] mostly has been concentrated around the silico-
Ether-rich fuel production [277] alumino-phosphate molecular sieve SAPO-34 with
Conversion of olefins to diesel [278] different modifications. Inui in Kyoto has been
Reactivated MTG catalyst [279] involved strongly in this research [283–289].
Hydrocarbon dewaxing utilizing a reacti- [280]
vated spent MTG zeolite catalyst
Characteristics of various zeolite catalysts for the
light olefin production ( like ZKU-4, Ca-ZSM-5,
28 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

Fe-silicate, SAPO-34 and Ni-SAPO-34) were whereas aromatic compounds were not formed at
investigated and summarized (see Fig. 8). all (see Fig. 8).
ZKU-4 zeolite has an offretite–erionite SAPO-34 has a chabazite structure, but much
intergrowth structure, and is regarded as a kind weaker acidity than chabazite, and exhibits activity
of small-pore zeolite. While narrow pore zeolites for methanol to light olefins conversion at a
such as chabazite and ZSM-34 were effective for medium range temperature of around 450°C. Ni
the methanol to hydrocarbons reaction, these two containing SAPO-34 produced ethylene with a
catalysts had short lifetimes due to rapid coke selectivity of 90 at 100% methanol conversion
deposition. Inui and coworkers have considerably without significant deactivation. This indicates that
improved ZSM-34 type catalysts in both olefin proper acidity control prevents the formation of
selectivity and catalyst lifetime by arranging crys- aromatics and allows highly selective conversion
talline size and uniformity and adding quinoline of methanol to light olefins [283,284]. The Ni
in the feed gas, however, the coke formation could containing SAPO-34 catalysts were prepared by
not be prevented completely [284]. Medium pore adding nickel nitrate to the solution for mixed gel
size zeolite ZSM-5 had a too strong acidity to preparation according to the rapid crystallization
produce light olefins with high selectivity [283]. method (4 h). By controlling the uniformity of the
H-ZSM-5 catalysts, modified with basic materials gel mixture, Ni-SAPO-34 crystals with Si/Ni ratios
( like Ca) are effectively used at higher reaction of 100 and 40 were prepared having a sharp crystal
temperatures such as 600°C and yield propylene size distribution around 0.8–0.9 mm. As shown in
as the major product by obeying b-scission [284]. Fig. 9, with increasing of Ni-incorporation the acid
However, a certain amount of aromatics was strength measured by ammonia-TPD profiles
formed as well. A pentasil Fe-silicate yielded ethy- decreases not only for the strong acid sites but
lene and propylene almost exclusively, at an equiv- also for the weak acid sites [289].
alent ratio at lower temperatures around 300°C, The corresponding results for the methanol con-

Fig. 8. Comparison of hydrocarbon distribution from methanol conversion on various zeolite catalysts (reproduced by permission of
Prof. T. Inui, Kyoto University, Japan).
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 29

density could be controlled to be 3.0 mmol m−2


[289].
In the early period of the investigations of MTO
catalysts, small-pore zeolites such as chabazite,
erionite, ZSM-34 and offretite were selected as the
target of the catalysts for this purpose, because
pore openings consisting of eight T-membered
rings were regarded as just fitting the size of small
olefin molecules. Indeed, those small-pore zeolites
gave considerably higher selectivities to ethylene
and propylene, however, rapid deactivation of the
catalyst performance, as a result of coke formation
in the cavities involving the pore channels, could
not be avoided. In order to mitigate the coke
Fig. 9. Ammonia-TPD profiles for SAPO-34 and Ni-SAPO-34 deposition, the acid sites of those materials were
catalysts (reproduced by permission of Elsevier Science B.V., modified by basic materials, but those kinds of
Amsterdam). post-modifications were not successful as desired.
It turned out to be quite difficult to neutralize the
version to olefins on SAPO-34 and different acid sites uniformly and adjusting the proper acid
Ni-containing SAPO-34 molecular sieves are sum- strength for selective olefin formation [289].
marized in Fig. 10. Although Ni-SAPO-34 is a small-pore micro-
Ni-SAPO-34 having a Si/Al ratio of 40 exhibited porous material like chabazite, its acidity was
an extraordinary high ethylene selectivity as high considerably weaker than that of chabazite and
as 90% at total methanol conversion conditions even SAPO-34. Obviously, this property just fits
[289]. From the extrapolation of the relationship for exclusive ethylene formation from methanol
between the internal acid density of various via dimethyl ether, and the consecutive reactions
Ni-SAPO-34 catalysts prepared by different meth- do not occur. As the oligomerization of ethylene
ods and the ethylene selectivity at total methanol and aromatization cannot happen on this catalyst,
conversion, it was concluded that the ethylene no coke formation is observed. This provides not
selectivity could arrive at 100% if the internal acid only a large potential to develop a novel ethylene

Fig. 10. Catalytic performance of SAPO-34 and Ni-SAPO-34s in methanol conversion. Reactions conditions: 20% MeOH–80%
nitrogen, GHSV 2000 h−1, temperature: 450°C (reproduced by permission of Elsevier Science B.V., Amsterdam).
30 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

synthesis process, but also a guiding principle for


the synthesis of other highly shape selective cata-
lysts [287].
Pop et al. [290] investigated the influence of
different operating parameters on the product dis-
tribution of the MTO process in a fixed-bed labora-
tory scale reactor using SAPO-34 as catalyst. They
observed 90% total olefin selectivity and a 61%
ethylene selectivity (435°C and LHSV 1.95 h−1).
Recirculation of dimethyl ether and unreacted
methanol was applied to improve the selectivities
by about 1%. Due to the high catalyst deactivation
rate, the continuous recirculation of the catalyst
in a fluidized bed reactor–regenerator system is
advisable. The proposed reduced kinetic model
could satisfactorily correlate the data at 400°C.
Further data and kinetic parameter calculations
could validate the model versus different operating
conditions [290].
A quantitative description of the active sites in
the dehydrated acid catalyst H-SAPO-34 for the
Fig. 11. IR spectra of Brønsted-OH in H-SAPO-34 at 77 K
conversion of methanol to olefins has been before (curve a) and after (curve b) the adsorption of CO. It is
reported by Smith et al. [291]. Their neutron revealed, using a fitting procedure, that both bands are consti-
diffraction study of the deuterated form of tuted of three components (A, B and C ). The overall fitted
SAPO-34 (D-SAPO-34) revealed two different spectrum is reported on top of the Figure (see also Table 5,
reproduced by permission of J.C. Baltzer Science Publishers,
bridging hydroxyl Brønsted sites. These sites are
Amsterdam).
also identified by IR spectroscopy which distingu-
ishes their acid strength from the shift in O-H
frequencies following adsorption of CO at 77 K scopy were used to characterize Ni-SAPO-34
(see Table 5 and Fig. 11). molecular sieves to be applied in the MTO reac-
Both their powder refinement results of tion. It was established that the Ni ions were all
D-SAPO-34 and the IR study of CO adsorption in tetrahedral, framework sites, and the loosely
of H-SAPO-34 point to a nearly equal occupancy bound extra-framework protons were thought to
of the two acidic protons [291]. be the key components of the catalyst for the
X-ray absorption (near-edge and extended fine highly selective conversion of methanol to ethylene
structure) along with diffuse reflectance IR spectro- [292]. The reported catalytic test results demon-

Table 5
IR spectroscopic features of the Brønsted OH groups in H-SAPO-34 before and after CO adsorption (according to Ref. [291])a

OH group n (cm−1) Areab OH,CO n (cm−1) Areab


OH OH-CO
A 3630 7.8 A∞ 3343 57.6
B 3625 3.1 B∞ 3276 10.2
C 3601 10.6 C∞ 3411 26.7
Fit A + B + C — 21.5 Fit A+B+C — 94.5
Curve aa — 20.8 Curve ba — 97.1

aRefers to Fig. 11.


bIntegrated intensity of the adsorption bands.
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 31

strate a clear effect of the temperature on the The hydrothermal treatment of SAPO-34 at
product distribution of the methanol conversion temperatures in excess of 700°C for periods suffi-
over Ni-SAPO-34 (Al/Ni ratio of 58.7) after 2 h cient to destroy a large portion of their acid sites
on stream: temperature increase from 250 to 450°C while retaining at least 80% of their crystallinity,
increases the methanol conversion (from 90.5 to was found to result in a catalyst for converting
95.5%), whereas the selectivity to ethylene methanol to lower olefins having an increased
decreases dramatically from 94.5 to 60.5% [292]. catalyst lifetime, increased selectivity for light ole-
The catalytic activity and selectivity of fins formation and decreased selectivity for paraffin
as-synthesized and modified samples of SAPO-34 production compared with the untreated SAPO-34
and Me-APSO-34 (MeNCo, Ni) for the conver- starting material [294].
sion of methanol to olefins have been investigated The methanol conversion to light olefins was
by van Niekerk et al. [293]. The catalytic perfor- investigated using different SAPO type molecular
mance for the conversion of methanol to light sieves by Chen et al. [295]. At a temperature of
olefins of all the catalyst samples prepared was 500°C and a WHSV of 1.5 h−1 SAPO-34 showed
found to be closely related to the number of strong the highest selectivity towards formation of ethy-
acid sites present as determined through ammonia- lene and propylene (almost 90%), followed by
TPD. Mild steaming, encountered during deep- SAPO-18 (ca. 80%), SAPO-17 (around 72%) and
bed calcination, increased the relative crystallinity finally SAPO-5 (about 53%). The bridging
and the lifetime of SAPO-34 due to the formation Si-OH-Al groups in these SAPOs were monitored
of stronger acid sites probably on the external by means of diffuse reflectance IR spectroscopy
surface of the crystallites as well as the catalytic (DRIFT ), and the strength of the Brønsted acid
performance. Selectivities to light olefins were typi- sites was measured by ammonia-TPD resulting in
cal of those previously reported (up to 92 wt%) the following order of Brønsted acid site strength:
and essentially constant for all the samples investi- SAPO-5<SAPO-17<SAPO-18<SAPO-34, the
gated. The almost complete absence of C+ olefins last one having cages smaller than those of SAPO-5
5
was ascribed to the so-called ‘cage-effect’ which and SAPO-17. The deactivation rate decreased
imposes a restriction on the formation of those in the order SAPO-17>SAPO-34>SAPO-18>
compounds. Dilution of the methanol with water SAPO-5 [295]. Yang et al. [296 ] compared the
as opposed to nitrogen increased the catalyst utili- catalytic behavior of the small-, medium- and
zation value three-fold and reduced the rate of large-pore SAPO-type materials in the methanol
coke formation during reaction. Treatments such conversion to hydrocarbons. They concluded that
as steaming, silanization and poisoning of strong SAPO-34 favors the formation of lower olefins,
sites by ammonia all reduced the number of strong SAPO-11 the production of C and C hydro-
5 6
acid sites and, thus, reduced catalytic perfor- carbons and SAPO-5 the formation of C+ hydro-
7
mance [293]. carbons [296 ]. The selectivities towards lower
The selectivity and stability of the H-SAPO-34 olefins increased with increasing temperature over
catalyst for the methanol to olefins reaction was SAPO-34 and SAPO-5. The effect of contact time
much better than those of erionite and erionite– was not significant on SAPO-34 and SAPO-11
offretite intergrowth type zeolites as reported by under the reaction conditions studied (reaction
Liang et al. [249]. These authors found that after temperatures, 350–400°C; WHSV, 0.4–0.9 h−1).
55 regenerations, H-SAPO-34 retains its catalytic The selectivity towards lower olefins increased with
performance (under the reaction conditions of increasing silicon content of SAPO-11. Lowering
450°C, WHSV of 6.9 h−1 and methanol/water the partial pressure of methanol increased the
ratio of 20:80) for the methanol conversion, the selectivity towards lower olefins over SAPO-5
total hydrocarbon yield, the content of light olefins [296 ].
and the ethylene formation level in the hydro- The reaction products of the methanol conver-
carbon products at values of 100, 98.6, 91.7 and sion to olefins over H-SAPO-34, H-chabazite and
56.6%, respectively [249]. H-ZSM-5 was compared by Nawaz et al. [297]. A
32 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

strongly shape-selective eight-membered ring


microporous material was needed to suppress the
formation of aromatics, and thereby get the highest
olefins selectivity. An olefin yield slightly above
80 mol% at close to 100% conversion of methanol
could be obtained with the H-SAPO-34 catalyst
(reaction conditions: temperature 400°C, WHSV:
1 h−1, feed consisting of 40 mol% methanol ). Fast
deactivation was the main limitation in the MTO
reaction. The deactivation rate, which was much
smaller in the case of H-ZSM-5, appears partly
structure dependent, but a large difference in deac- Fig. 12. Methanol conversion over SAPO-17 at various partial
tivation rate between the isostructural catalysts pressures at 400°C.
H-SAPO-34 and H-chabazite was observed
(H-chabazite deactivated much faster than methanol more deeply than SAPO-34. This must
H-SAPO-34) due to the higher acid site content be attributed to their different pore structures.
of H-chabazite compared with H-SAPO-34 [297]. SAPO-17 contains partly the SAPO-5 structure
The same group investigated SAPO-17 and its that is similar to the mordenite structure. This
isostructural zeolite erionite with respect to the permits formation of fused-ring aromatics, and
catalytic behavior in the methanol to olefins reac- furthermore, it contains a one-dimensional pore
tion [298]. SAPO-17 proved superior to erionite structure and therefore, effective diffusion is very
both concerning the olefin selectivity and the cata- low [285]. These facts might be the reason for the
lyst lifetime. Ethylene constituted almost 70 mol% higher deactivation rate of SAPO-17 compared
and propylene almost 20 mol% of the product, with SAPO-34.
compared with a maximum of 40 mol% ethylene The activity, selectivity and coke formation on
and 30 mol% propylene with erionite (at 425°C SAPO-34 and H-ZSM-5 at 425°C in the conversion
reaction temperature and 0.5 h−1 WHSV ). of methanol to hydrocarbons have been studied
SAPO-17 was active for up to 20 h, whereas erio- using a microbalance reactor [299]. A selectivity
nite only for less than 2 h. Experiments with to light olefins of 91 mol% at 100% conversion of
SAPO-17 in the reaction temperature range of methanol into hydrocarbons was observed over
350–475°C showed an optimum behavior around SAPO-34, but the catalyst suffered from rapid
425°C. Outside this temperature, the lifetime and deactivation. H-ZSM-5 showed a lower activity,
selectivity towards lower olefins deteriorated lower selectivity to light olefins and a lower rate
rapidly. A deactivated sample of SAPO-17 could of deactivation. Co-feeding methanol and ethylene
be restored to full activity by calcination in air at or propylene showed that methanol is responsible
500–550°C [298]. A number of catalytic test for the formation of methane and ethylene, and
experiments at 400°C were performed on SAPO-17 that added ethylene and propylene participate in
where the partial pressure of methanol was varied. secondary reactions [299].
From the plots of methanol conversion versus time SAPO-18 and its MeAPO analogues (MeNMg,
on stream ( Fig. 12), it is apparent that the catalyst Zn), which are isostructural with AlPO -18, are
4
lifetime increased when the methanol pressure was structurally closely related to chabazite/SAPO-34.
lowered. The yield of hydrocarbons before regener- Those AEI compounds have been investigated in
ation became necessary was, however, essentially the methanol to olefins reaction, and SAPO-18
constant. This may suggest that the coke formation revealed catalytic properties comparable to
is mechanistically connected with the hydrocarbon SAPO-34: 47 wt% selectivity towards ethylene and
formation [298]. 40 wt% selectivity towards propylene at t
DME
In spite of the fact that SAPO-34 has a higher (breakthrough of dimethyl ether defined as the
acid strength than SAPO-17, the latter converts time span with more than 99% conversion of
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 33

methanol and dimethyl ether). The catalytic test methanol and the selectivity to light olefins
conditions were as follows: temperature 420°C, decreased as well, however, both of them were
WHSV of 1 h−1, feed: methanol diluted with 60% higher than those for H-ACDM. Besides, the
nitrogen [300]. MgAPO-18 and ZnAPO-18 results of the Mössbauer spectra revealed that
showed much higher selectivities to propylene and Fe O took part in the MTO reaction and was
2 3
butenes and shorter catalyst lifetimes than partly reduced to FeO [303].
SAPO-18. AlPO -18 was inactive as MTO catalyst. Finally, a fluid catalytic cracking process for the
4
The SAPO- and MeAPO-18 systems formed thin conversion of a hydrocarbonaceous feedstock has
plates of the same dimensions (crystal size been disclosed. The process results in an increased
0.1–2.0 mm) which was fortunate for the study and yield of lower olefins, and methanol is typically
comparison of catalytic activity, since diffusion co-processed with the feedstock to obtain a proper
induced effects should be equal [300]. feed dispersion prior to injection into the FCC
MeAPSO molecular sieves (MeNTi, Fe, Zn, reactor. During cracking of the feedstock,
Mg, Mn, Co and others) were patented by UOP co-cracking of the methanol dispersant to lower
for a fluidized-bed-based methanol to olefins reac- olefins occurs, resulting in an increased yield of
tion, providing mainly high propylene selectivities light alkenes, typically ethylene and propylene.
(up to 53 wt%) at 480°C and a WHSV of around Suitably, an amount of methanol of up to 10 wt%
1.85 h−1 using a methanol/nitrogen feed consisting on feed is used, preferably 3.5–7.5 wt% on feed.
of about 78.5 mol% nitrogen [301]. The temperature employed in the FCC unit is
greater than 400°C, suitably from 500 to 800°C.
4.2. Composite materials The catalyst comprises a zeolite with a pore diame-
ter of 0.3–0.7 nm commonly used for FCC, for
Only limited and uncomplete information has instance ZSM-5. The zeolite catalyst is optionally
been available concerning the published material present as a physical mixture with a second sepa-
related to the application of composite catalysts rate zeolite with a pore diameter above 0.7 nm,
for the methanol-to-hydrocarbons process technol- for instance a physical mixture of ZSM-5 and
ogy. A Soviet high-silica zeolite TsVM (pentasil- zeolite Y [304].
type structure) was modified by Rostanin et al.
[302] with respect to the removal of undercrystal-
lized silicate fragments present at the surface of 5. Catalyst deactivation, coke formation and
the zeolite crystals. This catalyst was used for the regeneration
manufacture of high-octane gasoline at 370–380°C
applying a methanol–water feedstock with various The porous structure of large-pore molecular
compositions and recirculation of the hydro- sieves can easily host carbonaceous residues which
carbon gases. usually cause catalyst deactivation. This loss of
The surface acidity, surface structure and metha- activity by coke formation is rather rapid with
nol to olefins catalytic activity of Fe O /ACDM mordenite [116 ] due to the blockage of the uniform
2 3
zeolite catalysts were monitored by Qinghua et al. non-interconnecting channels by coke. Zeolites X
[303], however, without describing the ACDM and Y can initially accommodate some coke in the
zeolite system. Their results showed that calcina- supercages without blockage of the pore structure,
tion treatment did not influence the crystallinity whereas SAPO-5, MAPO-5 and MeAPO-5 (one-
of the H-ACDM zeolite, and impregnation treat- dimensional pore structure) deactivate in a similar
ment did not destroy the supporting zeolite. With manner as mordenite [11]. Swabb and Gates
increasing Fe content and higher calcination tem- studied the deactivation of H-mordenite in the
peratures, the number of acid sites and the amount conversion of methanol in the range of 100–240°C
ratio of Brønsted/Lewis sites on the zeolite surface [38]. Olefin polymerization occurred on the strong
decreased as shown by ammonia-TPD and acid sites, and these oligomers were strongly
pyridine-IR. In addition, the conversion rate of adsorbed. Traces of lower olefins were only
34 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

detected at 240°C. The deactivation of H- pronounced for molecular sieves with one-dimen-
mordenite seemed to be more related to coverage sional non-interconnecting channels [11].
of acid sites than to pore-blockage [11]. A gradientless method has been used to study
H-mordenite and H-ZSM-5 catalysts used in the the deactivation of high-silica zeolite ZVM for the
conversion of 13C-enriched methanol (90%) were synthesis of hydrocarbons from methanol [306 ].
investigated by in situ 13C CP MAS NMR spectro- It was shown that olefins were responsible for the
scopy in order to monitor the formation of carbo- coke formation and that the coke formed affected
naceous residues [305]. Depending on the zeolite, the rates of different reaction pathways for chemi-
differences in the nature of the carbonaceous resi- cally converting methanol and the reaction pro-
dues were detected: organic residues on H-ZSM-5 ducts in different ways. Furthermore, the authors
showed a broad aliphatic product distribution, and demonstrated that the conversion of methanol to
the aromatic–olefinic region revealed the presence hydrocarbons proceeds in a quasi steady-state pro-
of BTX compounds. However, a different situation cess with respect to the state of the catalyst chang-
was observed for H-mordenite: a much narrower ing throughout the course of the deactivation. A
distribution of paraffinic products was encoun- kinetic model was proposed for the deactivation
tered, and only propylene, propane and n-butane of the catalysts [306 ].
were properly detected. By contrast, the broader Review papers exist dealing with the coke forma-
aromatic region indicates the formation of a great tion and regeneration of medium-pore type zeolites
variety of aromatic compounds, including fused (mainly ZSM-5) applied in the methanol conver-
aromatic ring systems. The absence of palpable sion reaction [307,308], and a few highlights are
amounts of isoparaffins and C -aliphatics for summarized in the following:
5
H-mordenite results from its large pore size which During the MTG process the ZSM-5 catalyst
renders easier the conversion of lower olefins into becomes less effective due to three forms of
aromatics by mechanisms resembling conjunct deactivation:
polymerization. The formation of fused ring aro- (1) the deposition of a carbonaceous residue on
matics is also possible in the mordenite structure and in the ZSM-5 catalyst;
and explains the broader aromatic resonance which (2) an irreversible loss of activity due to the effect
has been observed [305]. of steam on the zeolite structure;
The rate of coke formation is lower when dea- (3) during regeneration, when coke is removed by
luminated mordenite is exchanged with Ba cations, oxidation, the high temperatures involved may
according to Sawa et al. [119]. After 8 h on stream affect the zeolite structure.
the amount of coke deposited per mol of reacted These processes can be controlled by minimizing
methanol diminished as the degree of Ba exchange the regeneration temperature, the reaction temper-
increased. This was explained by the weakening of ature and partial pressure of steam and adjusting
the Brønsted acid sites due to the Ba exchange [11]. operating procedures to minimize coke formation.
Froment et al. observed that on dealuminated The total number of cycles before a new ZSM-5
mordenite the selectivity changed with time on catalyst charge is required for the MTG process
stream. The yield of light olefins increased, while has not yet been reported, but catalyst life appears
the yields of paraffins and aromatics dropped, even to be longer than 2 years, the period once antici-
when the methanol conversion was maintained at pated [307].
100% (see Fig. 13). In a review on the influence of the pore structure
Coke would already be formed at early stages on the coking and deactivation of zeolites, Guisnet
of the methanol conversion and would cover the and Magnoux [308] concluded that both the rate
strong acid sites, thus reducing the conversion of of coke formation and the nature and/or distribu-
olefins to paraffins and aromatics. In conclusion, tion of the coke components depend on the pore
coke deposition can vary the selectivity on large- structure. The rate of coke formation also depends
pore molecular sieves by coverage of acid sites and on the operation conditions, of course, although
pore blockage. The effect of blockage is more sometimes it is difficult to discriminate between
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 35

Fig. 13. Methanol conversion into hydrocarbons on mordenite ( Temperature: 480°C, Si/Al=80, feed: 30–70 wt% methanol/water)
(reproduced by permission of The Royal Society of Chemistry, Cambridge).

the effects of the pore structure and the acidity its acid sites. Time-on-stream experiments revealed
[308]. that ZSM-5, although considered to be highly
One of the major advantages of ZSM-5 when resistant to deactivation by coke formation, could
compared with other zeolites is its high resistance also deactivate significantly due to the formation
to deactivation by coke formation [11]. It was of substantial amounts of coke. This was rational-
early recognized that intracrystalline coke forma- ized by either a limited access of the reactant by
tion on zeolites is a shape-selective reaction, blockage of the pore openings or by formation of
directly related to the pore structure [309,310]. By non-desorbable polymerization products inside the
comparing coke formation on ZSM-5 and morde- pores [11,314].
nite in the methanol conversion it was clearly Ione et al. [315] studied the effect of the Si/Al
demonstrated that the low selectivity towards coke ratio on the coke formation on ZSM-5 type zeolites
formation on ZSM-5 must arise from structural in the methanol reaction. An increase in the pro-
constraints on the reaction of the same intermedi- duction of aromatics, which, in turn resulted in a
ates of coke [11]. Dejaifve et al. [116 ] showed that faster deactivation, was observed with decreasing
the deposition of carbonaceous residues on ZSM-5 Si/Al ratio. This was confirmed by Bibby et al.
occurs essentially on its external surface and that [316 ] who noticed that the overall rate of coke
the active sites are not covered by coke. This formation was related to the aluminum content of
finding was supported by Schulz et al. [311] and ZSM-5, but not the initial rate of coke formation
Slinkin et al. [312] who studied the oligomerization and the ultimate amount of coke formed. It was
of olefins on ZSM-5 and proposed that the coke also shown that coke formation continued, even
precursors were formed inside the catalysts, but after complete deactivation, due to thermal crack-
have a high mobility, due to the low density of ing of methanol. Suzuki et al. [145] monitored the
acid sites, so that they are readily desorbed to effect of the crystallization time on the properties
form coke on the outer surface. In this scheme, of ZSM-5. The Al content at the external surface
the density of the acid sites inside the channels increased with crystallization time, while the cata-
may be a more important factor for the resistance lyst stability in the methanol conversion
to coke formation than the shape selectivity [11]. decreased [11].
Magnoux et al. [313] attribute the low coking rate Bibby et al. [316,317] presented experimental
on ZSM-5 to a large extent to the low density of evidence that part of the coke formed during the
36 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

methanol conversion must be inside the channel of methane to be symptomatic for coke laydown.
system. They found a linear relationship between The main mechanism of methane formation from
the sorption capacity (expressed as void volume) methanol would be a non-ionic methylation of the
and the coke formation (expressed as a volume coke deposits, followed by elimination of meth-
fraction of the zeolite volume). McLellan et al. ane [11].
[318] monitored the acidity of the ZSM-5 catalyst In situ NMR spectroscopy has been applied to
as a function of the coke content and concluded investigate reaction pathways leading to coke for-
that coke formation occurs initially at the channel mation during the methanol conversion over
intersections. In a second stage, external coke was ZSM-5 by Anderson and Klinowski [72,73]. These
formed and topological blocking of the zeolite authors recorded 13C NMR spectra of zeolite
channels was observed. This was confirmed by samples in sealed tubes after heating in methanol
means of an XPS study of the coke formation to successively higher temperatures. The spectra
during the methanol conversion on ZSM-5 [319]. of adsorbed products showed the initial hydro-
Internal coke formation occurred predominantly carbon species to be aliphatic in nature, with
during the conversion of methanol until the cata- aromatics first detected at 300°C [72,73,307].
lyst was essentially deactivated. Only then, after The influence of water formed during the metha-
8 wt% coke had already been deposited inside the nol conversion on the coke formation chemistry
catalyst, was external coke formed, probably due has not yet been clearly established. Water at high
to thermal cracking of methanol [11]. This was temperatures can modify the acidity of high-silica
confirmed by Behrsing et al. [320] who investigated zeolites by hydrolyzing aluminum from the zeolite
carbonaceous deposits on ZSM-5 during methanol lattice, which may in turn affect the coking beha-
conversion at 460°C with TEM and observed vior of the zeolite [307].
external coke at approximately 6 wt% coke. Pentasil type zeolite catalysts with different Si/Al
According to Schulz et al. [311], the initial coke ratios and carbonized during the methanol conver-
formation during methanol conversion probably sion at 380–420°C have been studied by ESR
proceeds relatively slowly via carbenium ions, ole- spectroscopy [324]. The appearance of condensed
fins or aromatics as precursors. Further coke for- coke structures, which provided a narrow ESR
mation is fast and results mainly from direct signal, took place at low temperatures (about
reaction of methanol with the coke [11]. 400°C ) and was determined by the time of coke
The coke formation and deactivation on ZSM-5, formation and the distribution density of active
ZSM-11 and ZSM-48 was followed by Froment sites in the pentasil type zeolite.
et al. [321] using a microbalance in conjunction Pop et al. [325] investigated the kinetics of the
with on-line GC. In this way it was possible to coke deposits formation and their influence on the
obtain evidence about the effect of coke on both product selectivities by running experiments in a
the main and the coking reaction. For ZSM-5 and demonstration scale fluidized bed reactor for the
ZSM-11, but not for ZSM-48, the effect of coke MTG process. New reaction indices were advanced
was more pronounced on its own formation than for on-line monitoring of the catalyst activity.
on the rate of the main reaction [11]. Computer simulated deactivation of a ZSM-5
Chen et al. [322] found that the yield of methane zeolite during methanol conversion to gasoline
was closely related to the extent of coking in their with randomly distributed catalytic sites was per-
investigation of the deactivation of modified formed by Nelson and Bibby [326 ]. Data from
ZSM-5 catalysts in the MTO reaction at high this ‘random lattice’ simulation was in qualitative
temperature. Since methane and the C+ fraction agreement with available experimental results,
5
gradually increased with time on stream, it was whereas simulated deactivation with periodically
suggested that the coke deposition on the acid sites distributed catalytic sites was not. The main
enhanced the secondary cracking reaction through difference between the ‘random lattice’ simulation
a non-ionic mechanism [11]. Langner [323] and data and experimental results was that the number
Schulz et al. [311] also believe significant amounts of accessible intersections in the simulation was
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 37

larger than that expected from experimental acces- occupy strongly acidic sites [328]. According to
sible volumes in deactivated zeolites. This differ- Benito et al. [329], coke deposition takes place
ence may be explained if the ‘coke molecule’ almost exclusively in the internal channels of
covering an active site ‘grows’ to exclude a volume H-ZSM-5 (deposition of paraffins, oligomers and
greater than that of a single intersection (see alkylated monoaromatics) and in the intersections
Fig. 14). between these channels (deposition of alkylated
The influence of the carrier gas on the deactiva- mono- and biaromatics) until the catalyst becomes
tion rate and on the nature of coke on an H-ZSM-5 totally deactivated for hydrocarbon production.
catalyst during methanol conversion was studied The high branching of the H-ZSM-5 zeolite micro-
by Bauer et al. [327]. Recycling hydrogen-contain- porous structure explains why, in spite of the
ing light gases was found to prolong the time on progressive blockage of internal channels, active
stream and to enable partial reactivation of spent sites remain accessible from lateral channels and
zeolite catalysts. In spite of significant dealumina- deactivation by pore blockage is avoided. This
tion of H-ZSM-5 during methanol conversion at interpretation coincides with the kinetic results of
temperatures of about 415°C no loss of activity deactivation, which follow an equation dependent
was observed over several reaction cycles indicat- on the concentration of the reaction components,
ing that the MTG reaction occurs on relatively and corresponds to active-site blockage mecha-
few sites [327]. nisms. Furthermore, the acidity as well as the
The nature, deposition and the role of coke activity decreased almost linearly with the coke
characteristics in the regeneration of H-ZSM-5 as content. It was proven that deposited coke is very
catalyst for the MTG process have been investi- unstable [329] and undergoes severe aging (with a
gated [328–330]. The coke deposited on the H/C ratio decrease) during the thermal treatment
H-ZSM-5 catalyst in the MTG process in the at regeneration temperature. Coke instability will
250–450°C range was basically soluble in pyridine have a strong influence on regeneration kinetic
and had a structure constituted by methyl-substi- studies and on catalyst regeneration in the reactor.
tuted polycyclic aromatics. The molecular weight On the basis of their results, Ortega et al. [330]
of the coke components and their C/H ratio concluded that aging treatment by sweeping with
increased with time on stream and temperature. an inert gas was needed in both cases in order to
The acidity deterioration occurred simultaneously obtain reproducible and well-controlled results
with the increase in content of coke deposited (up under conditions in which coke was equilibrated.
to 1.6 wt% with respect to the catalyst at 350°C ) In spite of the high selectivity for linear light
and was attributable to the irreversible occupation hydrocarbons, the small-pore molecular sieves gen-
of acid sites. The coke components preferably erally suffer from relatively rapid deactivation
during the methanol conversion reaction [11]. The
large cavities are responsible for this rapid deacti-
vation. For example, the fast deactivation of zeolite
T (erionite–offretite intergrowth) was attributed to
the blockage of the cavities by coke at 400°C and
by methoxy groups at 200–300°C. Aromatics and
heavy branched compounds can be formed inside
the large cages at temperatures above 300°C. These
molecules cannot diffuse through the porous struc-
ture because their kinetic diameter is larger than
Fig. 14. Schematic representation of the volume (black area) in the pore-opening size of the small-pore molecular
zeolite ZSM-5 excluded by a ‘coke molecule’: (a) excluded sieves. Thus, they remain inside the large cages
volume is one intersection; and (b) excluded volume is more
than one intersection — assumes coke growing into adjacent where they can form carbonaceous deposits block-
channels (reproduced by permission of Elsevier Science B.V., ing the pore openings and preventing the access
Amsterdam). of reagent molecules to the active sites. The rate
38 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

of deactivation depends on the acidity: the higher Subsequently, the pores are blocked and the
the catalyst acid strength, the faster the deactiva- concentration of acid sites decays abruptly. The
tion by coke formation [11]. To minimize the methanol conversion then drops from 100% to a
deactivation of HT zeolite by coke formation, Ni- final value, which depends on the temperature and
and Pt-exchanged samples have been used, and the water concentration in the feed. In this final
the reaction was followed in the presence of stage the principal products are dimethyl ether,
hydrogen. The deactivation of Pt-HT was much methane and small amounts of light olefins (see
slower than that of HT but only paraffinic hydro- Fig. 15) which are formed by reactions requiring
carbons were produced [331]. The H-form of only weak acid sites [11,247].
chabazite, erionite and T-type zeolites deactivate Controversial debates have been conducted in
faster than dealuminated H-chabazite or connection with the irreversible deactivation of
H-SAPO-34 and H-SAPO-17 [11]. clinoptilolite samples. Sakoh et al. [236 ] claim that
As mentioned earlier, deactivation by coke may the activity of clinoptilolite can be completely
be due to both covering of acid sites and blockage restored by recalcination in air, Hutchings et al.
of the pore structure [11, 332–334]. The intercon- [235] observed the opposite in their study of
necting three-dimensional network of pores with deactivation and regeneration of clinoptilolite cat-
supercages existing in small-pore molecular sieves alysts modified by either ammonium ion exchange
provides room for accommodating some coke or hydrochloric acid treatment [11]. Gubisch and
without immediately blocking the pores [332]. It Bandermann monitored the behavior of HCl-
has been suggested that the deactivation by coke exchanged zeolite Na–T as a function of the
and its effect on the product distribution depend number of regeneration cycles [225]. Regeneration
on the way the coke is deposited on the catalyst was carried out between 420 and 520°C, and they
[247]. Initially, aromatics and branched isomers found that the catalyst activity decreased slowly
formed inside the cavities are adsorbed irreversibly from cycle to cycle, regardless of the regeneration
on the strong acid sites. As a consequence, the temperature. After 16 regeneration cycles the con-
concentration of strong acid sites decreases and so version of methanol dropped from 100 to 80% [11].
does the olefin conversion to paraffins, as Liang et al. [249] regenerated SAPO-34, erionite
illustrated in Fig. 15. and erionite–offretite zeolite samples at 530°C in

Fig. 15. Methanol conversion into hydrocarbons on SAPO-34 (temperature: 480°C, Si/Al=0.15, feed: 30–70 wt% methanol/water)
(reproduced by permission of Elsevier Science B.V., Amsterdam).
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 39

air after methanol conversion at 450°C and 5 h−1 while the conversion of methanol changed with
WHSV. With SAPO-34 the initial conversion of coke formation significantly, which was most likely
methanol after 55 regeneration cycles was still caused by the decrease in the diffusion rate due to
100%, while the total hydrocarbon yield and the coke formation on the catalyst. Furthermore, the
level of light olefins in the hydrocarbon mixture authors observed that the MTO reaction was
were constant, too. The offretite–erionite type clearly influenced by the methanol diffusion.
samples suffered from an appreciable deterioration Finally, the operation conditions are very impor-
after repeated regenerations [11]. tant with respect to the stability of small-pore
Selective deactivation was demonstrated for the molecular sieves. Marchi and Froment [247]
MTO reaction over SAPO-34 by Chen et al. [335]. demonstrated that it is possible to suppress the
The reason for this selective deactivation can be steps that involve coke formation on SAPO-34 by
quite complex. The coke deposition resulted in increasing the temperature and the water content
changes not only in intrinsic selectivity but also in in the feed. Obviously, water weakens the strong
shape selectivity. The coke formation during MTO acid sites responsible for hydrocarbon trans-
was found to be a transition-state shape selective fer reactions. With H-T zeolite, Gubisch and
reaction [335]. The effect of coke formation during Bandermann [225] found that increasing methanol
the methanol to olefins reaction over SAPO-34 at partial pressure and WHSV favor the yield of coke
425°C was further studied by the same group, due to an increasing influence of hydrogen transfer
which made the following important observations. reactions [11].
(1) Methanol was converted to dimethyl ether
both externally and internally over SAPO-34. The
dimethyl ether formation at the external surface 6. Commercial implications
was not significant, probably due to the relatively
weak external acidity of the catalyst. The forma- Mobil’s methanol to gasoline (MTG) process
tion of dimethyl ether at the external surface gave has attracted worldwide attention. Its discovery
a slightly lower rate for olefins formation. The has been hailed as the first new route in more than
methanol to dimethyl ether ratio found from the 40 years for the production of gasoline from coal.
experiments was far from equilibrium. (2) The commercial development effort has been cen-
The coke recorded as the total mass increase in tered on two versions of reactor design, the fixed-
the ‘tapered element oscillating microbalance bed and the fluidized-bed concepts. The control
( TEOM )’ reactor, could be divided into active and dissipation of the heat generated by the exo-
coke formed from oxygenates and inactive coke thermic conversion reaction is the major concern
formed from olefins. The active coke promoted in reactor design [338,339]. The first commercial
the conversion to olefins and could in fact represent fixed-bed unit, located in New Zealand, used steam
the true surface intermediate of the reaction. (3) generation and reactor effluent gas recycle to con-
Coke reduced the dimethyl ether diffusivity, which trol the reactor temperature. The fluidized-bed
enhanced the formation of olefins. (4) The coke version has been operated in a 4000 ton year−1
formation enhanced the ethylene formation. The scale in Wesseling (Germany). The advantage of
ethylene to propylene ratio increased with that pilot plant was the desired selectivity of hydro-
intracrystalline coke content, regardless of the carbons in the product mixture (cf. Table 3).
nature of the coke [336 ]. Besides producing high quality gasoline (MTG
Chen et al. [337] continued their work by study- operation) process conditions can be varied to
ing the adsorption of the reactant, the reaction, produce olefins (MTO operation), which can be
the desorption of the products and the coke forma- converted to gasoline and diesel. The fluid-bed
tion as a function of the coke content in the technology provided all advantages in terms of
TEOM reactor under conditions of the MTO increased product yield, better quality and a very
reaction over SAPO-34. The adsorption of metha- efficient heat recovery. Depending on the world
nol changed only slightly with coke formation, marked situation (price, demand, etc.) the fluid-
40 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

ized-bed technology is available to produce liquid SAPO-34 a significant advantage over other types
fuels via methanol [340]. of catalyst systems.
Nowak et al. [341] and Martin et al. [342,343] The need to remove the high exothermic heat
combined the exothermic MTO reaction with the of the MTO reaction as well as the need for
endothermic steam cracking of hydrocarbons by frequent regeneration led to a fluidized-bed reactor
adjusting the operation conditions in such a way and regenerator design. This design also required
that the overall process was nearly thermoneutral. development of a fluidized-bed MTO catalyst with
This was favorable from an engineering point of the strength and integrity to last in fluidized-bed
view since the problems associated with heat conditions. UOP has commercially manufactured
removal in MTO and heat generation in steam the MTO SAPO-34 catalyst that has shown the
cracking were avoided. The appropriate temper- type of attrition resistance and stability that
ature for a balanced operation of this coupled enables it to handle multiple regenerations and
methanol–hydrocarbon cracking (CMHC ) is in fluidized-bed conditions. Multiple regeneration
the range of 600–700°C using ZSM-5 type zeo- studies in a small-scale as well as in a larger
lites [344]. reactor-regeneration demonstration unit (process-
It is likely that future commercialization of the ing methanol feed of 0.5 MTD) have been per-
methanol-to-olefins (MTO) process will take place formed by UOP–Hydro. Fig. 16 shows the results
in a fluidized-bed reactor for many of the same of 50 h of continuous methanol-to-olefins opera-
reasons which encouraged fluidized-bed MTG tion performed in the large demonstration plant
development, including better temperature control and using SAPO-34.
and constant product composition. The olefins The spent catalyst is circulated to the regenera-
tor, where coke is burned off, and then returned
produced by this process can be readily converted
to the reactor to achieve a steady state. Steam is
to gasoline, distillate and/or aviation fuels by
also generated in the regenerator to remove the
commercially available technologies ( like Mobil’s
exothermic heat from coke burning. The
MOGD process) [345].
UOP–Hydro MTO process can increase total ole-
UOP and Norsk Hydro introduced a selective
fins by more than 60% on a per pass basis and
and economical route for converting natural gas
offers a flexible ethylene/propylene ratio between
to olefins, the GTO process. The first step in this
0.75:1 and 1.5:1 [347]. Finally, since the MTO
process is natural gas conversion to methanol
process for the production of ethylene and propyl-
followed by the UOP–Hydro MTO process using ene requires a large quantity of methanol feed, the
UOP’s unique SAPO-34 catalyst. The primary methanol production should be at the same site as
products are ethylene and propylene, two large- the MTO process. Both the UOP–Hydro MTO
volume base petrochemicals [3,346,347]. SAPO-34 process and the SAPO-34 catalyst are currently
was found to be catalytically very selective for available for licence from UOP and Hydro [3,346 ].
methanol conversion to ethylene and propylene.
The UOP–Hydro MTO process has the advantage
of high selectivity to ethylene due to this catalyst. 7. Conclusions
Ethylene has been shown to be the primary product
of this reaction. Ethylene subsequently oligo- Over the last few decades, methanol-to-hydro-
merizes to higher compounds, which are thermo- carbons technologies have been the focus for a
dynamically favored but are produced at substan- large number of researchers dealing with the
tially lower levels with the SAPO-34 catalysts than upgrading of natural resources beneficial both for
with the ZSM-5 catalysts. Although these com- the petrochemistry and fuel industries. The strong
pounds may exist within the SAPO-34 cavities, emphasis on the R&D efforts resulted in the build-
only small linear olefins and paraffins readily pass ing of an MTG plant in New Zealand by Mobil
through the <4 Å diameter pores of the molecular and a licence offer by UOP–Hydro for the MTO
sieve. This high selectivity to ethylene gives process. However, the changing marked situation
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 41

Fig. 16. Continuous test results of the UOP/Hydro SAPO-34 catalyst in the MTO demonstration plant (reproduced by permission
of Elsevier Science B.V., Amsterdam).

worldwide with respect to the demand and price version of olefins to paraffins. The methods applied
for the final and/or intermediate products governs to reduce the concentration of strong acid sites in
the present and future development with respect zeolites/microporous materials are dealumination,
to the application of all parts of those technologies. cation exchange and isomorphous substitution of
Besides the process engineering aspects, a huge Al by polyvalent cations [11].
knowledge concerning the involved microporous SAPO molecular sieves, with their mild acidity,
materials has been gathered during this research, are very interesting alternatives to obtain high
especially with respect to ZSM-5 (MTG directed ) selectivities for light olefins. The conditions of
and SAPO-34 (MTO directed ) which is of benefit synthesis, dealumination and cation exchange seem
for the ‘zeolite’ community in general. While the to be very important for obtaining a catalyst with
FCC research promoted the insight into the field a good performance for methanol conversion into
of large-pore zeolites/microporous materials, the light olefins. Furthermore, the ratio of acid sites
methanol to hydrocarbons research contributed on the external surface to acid sites in the porous
strongly to the enhancement of the knowledge structure also plays an important role. The
directed towards the medium- and small-pore smaller the crystallite size, the higher this ratio.
molecular sieves. Finally, the presence of certain OH groups
It follows from the results reviewed that (n =3600 cm−1, which correspond to Al-OH
OH
medium-pore zeolites/microporous materials yield bonds) were found to be the most active ones for
generally C –C hydrocarbons, whereas small- the methanol conversion into hydrocarbons.
5 11
pore molecular sieves usually generate C –C Consequently, it is very important that the access
2 4
hydrocarbons. This may be related to diffusion to these acid sites is not blocked [11].
constraints and cavity dimensions. However, the The research directed towards methanol-to-
porous structure is not the only factor which must hydrocarbons is still in progress and new concepts
be taken into account in order to reach, for are under development and will contribute to an
example, a high selectivity to C –C olefins. High enhanced knowledge within this field, like the
2 4
selectivities for light paraffins, principally propane, application of ‘Zebedee’ (acronym for zeolites by
can be observed on small-pore zeolites. The reduc- evolutionary de novo design) to prepare catalysts
tion of the concentration of acid sites, responsible for the production of ethylene and propylene [348].
for hydrogen transfer reactions, decreases the con- The author is aware of the fact that not all
42 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

contributions published so far could be taken into [8] S. Yurchak, in: D.M. Bibby, C.D. Chang, R.F. Howe, S.
account in connection with the preparation of this Yurchak ( Eds.), Methane Conversion, Elsevier,
Amsterdam, 1988, p. 251.
review. However, he hopes that this paper can [9] J. Topp-Jørgensen, in: D.M. Bibby, C.D. Chang, R.F.
contribute to provide an overview within this topic Howe, S. Yurchak ( Eds.), Methane Conversion, Elsevier,
of heterogeneous catalysis, which plays an impor- Amsterdam, 1988, p. 293.
tant role from a technological and (at least for [10] I.J. Miller, in: D.M. Bibby, C.D. Chang, R.F. Howe, S.
certain countries) industrial point of view. Yurchak ( Eds.), Methane Conversion, Elsevier,
Amsterdam, 1988, p. 325.
[11] G.F. Froment, W.J.H. Dehertog, A.J. Marchi, A review
of the literature, Catalysis 9 (1992) 1.
8. Note added in proof [12] C.D. Chang, in: D.M. Bibby, C.D. Chang, R.F. Howe,
S. Yurchak ( Eds.), Methane Conversion, Elsevier,
Inui and Kong [349] have reported a reliable Amsterdam, 1988, p. 127.
[13] G.H. Hutchings, R. Hunter, Catal. Today 6 (1990) 279.
synthetic procedure for Ni-SAPO-34 crystals [14] J.P. van den Berg, J.P. Wolthuizen, J.H.C. van Hooff, in:
(0.85 mm in size) having the catalytic performance L.V. Rees ( Ed.), Proceedings 5th International Zeolite
of methanol to ethylene with a selectivity as high Conference (Naples), Heyden, London, 1980, p. 649.
as 88%. Goguen et al. [350] have revealed a [15] G.A. Olah, H. Doggweiler, J.D. Felberg, S. Frohlich,
hydrocarbon-pool mechanism in MT6 chemistry M.J. Grdina, R. Karpeles, T. Keumi, S. Inaba, W.M. Ip,
K. Lammertsma, G. Salem, T.C. Tabor, J. Am. Chem.
on H-ZSM-5 using a pulse–quench catalytic reac- Soc. 106 (1984) 2143.
tor monitored by in-situ NMR spectroscopy. [16 ] G.A. Olah, Pure Appl. Chem. 53 (1981) 201.
[17] G.A. Olah, H. Doggweiler, J.D. Felberg, J. Org. Chem.
49 (1984) 2112.
[18] G.A. Olah, H. Doggweiler, J.D. Felberg, J. Org. Chem.
Acknowledgements
49 (1984) 2116.
[19] R. Huisgen, Angew. Chem. 67 (1955) 439.
The author is indebted to Prof. Jens Weitkamp [20] V. Franzen, L. Fikentscher, Lieb. Annal. Chem. 617
( University of Stuttgart, Germany) for his kind (1957) 1.
encouragement during the preparation of this [21] R. Hunter, G.J. Hutchings, J.C.S. Chem. Comm.
review. Thanks are due to Tordis Whist for techni- (1985) 886.
[22] R. Hunter and G.J. Hutchings, J.C.S. Chem. Comm.
cal assistance in connection with the preparation (1985) 1643.
of the figures, and the relevant publishers/authors [23] C.S. Lee and M.M. Wu, J.C.S. Chem. Comm. (1985) 250.
are gratefully acknowledged for their permission [24] C.T.-W. Chu, C.D. Chang, J. Catal. 86 (1984) 297.
to reproduce the figures as cited in the text. [25] S.D. Hellring, C.D. Chang, 21st ACS State-of-the-Art
Symp., Methanol as a Raw Material for Fuels and
Chemicals, Marco Is., Florida, USA, June 1986.
[26 ] T.R. Forester, S.-T. Wong, R.F. Howe, J.C.S. Chem.
References Comm. (1986) 1611.
[27] K.-J. Chao, L.-J. Huarng, Proceedings 8th International
[1] S.L. Meisel, Chemtech. 1 (1988) 32. Congress on Catalysis (Berlin), Verlag Chemie,
[2] C.D. Chang, A.J. Silvestri, Chemtech. 10 (1987) 624. Weinheim, 1984, p. V-667.
[3] B.V. Vora, T.L. Marker, P.T. Barger, H.R. Nilsen, S. [28] T. Mole, J.A. Whiteside, J. Catal. 75 (1982) 284.
Kvisle, T. Fuglerud, in: M. de Pontes, R.L. Espinoza, [29] T. Mole, J. Catal. 84 (1983) 423.
C.P. Nicolaides, J.H. Scholz, M.S. Scurrell ( Eds.), Stud. [30] D. Farcasiu, J. Catal. 82 (1983) 252.
Surf. Sci. Catal., vol. 107, Elsevier, Amsterdam, 1997, [31] M.B. Sayed, R.P. Cooney, Aust. J. Chem. 35 (1982) 2483.
p. 87. [32] M.B. Sayed, R.A. Kydd, R.P. Cooney, J. Catal. 88
[4] C.D. Chang, Catal. Rev.-Sci. Engng 25 (1983) 1. (1984) 137.
[5] C.D. Chang, Catal. Rev.-Sci. Engng 26 (1984) 323. [33] M.B. Sayed, J.C.S. Faraday Trans. I 83 (1987) 1771.
[6 ] C.D. Chang, in: A. Holmen, K.-J. Jens, S. Kolboe (Eds.), [34] G.J. Hutchings, L. Jansen van Rensburg, W. Pickl, R.
Stud. Surf. Sci. Catal., vol. 61, Elsevier, Amsterdam, Hunter, J. Chem. Soc. Faraday Trans. I 84 (1988) 1311.
1991, p. 393. [35] G.J. Hutchings, F. Gottschalk, M.V.M. Hall, R. Hunter,
[7] A.A. Avidan, in: D.M. Bibby, C.D. Chang, R.F. Howe, J. Chem. Soc. Faraday Trans. I 83 (1987) 571.
S. Yurchak ( Eds.), Methane Conversion, Elsevier, [36 ] C.D. Chang, A.J. Silvestri, J. Catal. 47 (1977) 249.
Amsterdam, 1988, p. 307. [37] P.B. Venuto, P.S. Landis, Adv. Catal. 18 (1968) 259.
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 43

[38] F.A. Swabb, B.C. Gates, Ind. Engng Chem. Fundament. [67] F.X. Cormerais, G. Perot, F. Chevalier, M. Guisnet,
11 (1972) 540. J. Chem. Res. S (1980) 362.
[39] P. Salvador, W. Kladnig, J.C.S. Faraday Trans. I 73 [68] S. Kolboe, Acta Chem. Scand. A40 (1986) 711.
(1977) 1153. [69] S. Kolboe, in: L. Guczi, F. Solymosi, P. Tétényi ( Eds.),
[40] C.D. Chang, J. Catal. 69 (1981) 244. Proceedings of the 10th International Congress on
[41] C.D. Chang, C.T.-W. Chu, J. Catal. 74 (1982) 203. Catalysis, Akadémiai Kiadó, Budapest, 1993, p. 449.
[42] W. Kirmse, Carbene Chemistry, Academic Press, New [70] R.M. Dessau, J. Catal. 99 (1986) 111.
York, 1971. [71] V.B. Kazansky, N.D. Chuvylkin, I.N. Senchenya,
[43] G.A. Olah, G. Klopman, R.H. Schlosberg, J. Am. Chem. Kinetics Catal. 27 (1986) 1275.
Soc. 91 (1969) 3261. [72] M.W. Anderson, J. Klinowski, Nature 339 (1989) 200.
[44] J.H.C. van Hooff, J. Catal. 79 (1983) 242. [73] M.W. Anderson, J. Klinowski, J. Am. Chem. Soc. 112
[45] C.D. Chang, C.T.-W. Chu, J. Catal. 79 (1983) 244. (1990) 10.
[46 ] D.V. Dass, R.W. Martin, A.L. Odell, G.W. Quinn in [74] E.J. Munson, N.D. Lazo, M.E. Moellenhoff, J.F. Haw,
D.M. Bibby, C.D. Chang, R.F. Howe, S. Yurchak ( Eds.), J. Am. Chem. Soc. 113 (1991) 2783.
Methane Conversion, Elsevier, Amsterdam, 1988, p. 177. [75] E.G. Derouane, J.B. Nagy, P. Dejaifve, J.H.C. van
[47] G.A. Olah, G.K. Surya Prakash, R.W. Ellis, J.A. Olah, Hooff, B.P. Spekman, J.C. Vedrine, C. Naccache,
J.C.S. Chem. Comm. (1986) 9. J. Catal. 53 (1978) 40.
[48] W. Drenth, W.T.M. Andriessen, F.B. van Duijneveldt, [76 ] F. Salehirad, M.W. Anderson, J. Catal. 164 (1996) 301.
J. Mol. Catal. 21 (1983) 291. [77] M.W. Anderson, B. Sulikowski, P.J. Barrie, J. Klinowski,
[49] J. Nováková, L. Kubelková, Z. Dolejšek, J. Catal. 108 J. Phys. Chem. 94 (1990) 2730.
(1987) 208. [78] Y. Xu, C.P. Grey, J.M. Thomas, A.K. Cheetham, Catal.
[50] G.J. Hutchings, R. Hunter, W. Pickl, L. Jansen van Lett. 4 (1990) 251.
Rensburg, in: D.M. Bibby, C.D. Chang, R.F. Howe, S. [79] E.J. Munson, A.A. Kheir, N.D. Lazo, J.F. Haw, J. Phys.
Yurchak (Eds.), Methane Conversion, Elsevier, Chem. 96 (1992) 7740.
Amsterdam, 1988, p. 183. [80] M.W. Anderson, M.L. Occelli, J. Klinowski, J. Phys.
[51] Y. Ono, T. Mori, J.C.S. Faraday Trans. I 77 (1981) 2209. Chem. 96 (1992) 388.
[52] J.B. Nagy, J.P. Gilson, E.G. Derouane, J. Mol. Catal. 5 [81] V. Bosáček, J. Phys. Chem. 97 (1993) 10732.
(1979) 393. [82] G. Mirth, J.A. Lercher, M.W. Anderson, J. Klinowski,
[53] D. Kagi, J. Catal. 69 (1981) 242. J. Chem. Soc. Faraday Trans. I 86 (1990) 3039.
[54] S. Kolboe, in: D.M. Bibby, C.D. Chang, R.F. Howe, S. [83] D.V. Dass, R.W. Martin, A.L. Odell, J. Catal. 108
Yurchak (Eds.), Methane Conversion, Elsevier, (1987) 153.
Amsterdam, 1988, p. 189. [84] T.R. Forester, R.F. Howe, J. Am. Chem. Soc. 109
[55] T. Mole, in: D.M. Bibby, C.D. Chang, R.F. Howe, S. (1987) 5076.
Yurchak (Eds.), Methane Conversion, Elsevier, [85] S. Ceckiewicz, J. Chem. Soc. Faraday Trans. I 77
Amsterdam, 1988, p. 145. (1981) 269.
[56 ] R.D. Smith, J.H. Futrell, Chem. Phys. Lett. 41 (1976) 64. [86 ] G.J. Hutchings, P. Johnston, Appl. Catal. 67 (1990) L5.
[57] C.D. Chang, S.D. Hellring, J.A. Pearson, J. Catal. 115 [87] G.J. Hutchings, R. Hunter, P. Johnston, L. Jansen van
(1989) 282. Rensburg, J. Catal. 142 (1993) 602.
[58] W. Zatorski, S. Krzyzanowski, Acta Phys. Chem. 29 [88] G.J. Hutchings, D.F. Lee, M. Lynch, Appl. Catal. A106
(1978) 347. (1993) 115.
[59] J.K.A. Clarke, R. Darcy, B.F. Hegarty, E. O’Donoghue, [89] J. Nováková, L. Kubelková, Z. Dolejšek, J. Mol. Cat.
V. Amir-Ebrahimi, J.J. Rooney, J. Chem. Soc. Chem. 45 (1988) 365.
Comm. (1986) 425. [90] S. Kolboe, Acta Chem. Scand. A42 (1988) 185.
[60] H. Choukroun, D. Brunel, A. Germain, J. Chem. Soc. [91] R.F. Howe, in: D.M. Bibby, C.D. Chang, R.F. Howe, S.
Chem. Comm. (1986) 6. Yurchak ( Eds.), Methane Conversion, Elsevier,
[61] R. Hunter, G.J. Hutchings, W. Pickl, J. Chem. Soc. Amsterdam, 1988, p. 157.
Chem. Comm. (1987) 843. [92] V. Bosáček, Z. Tvaružková, Coll. Czech. Chem. Comm.
[62] R. Hunter, G.J. Hutchings, W. Pickl, J. Chem. Soc. 36 (1991) 551.
Chem. Comm. (1987) 1369. [93] E.G. Derouane, J.P. Gilson, J.B. Nagy, Zeolites 2
[63] G.J. Hutchings, in: A. Holmen, K.-J. Jens, S. Kolboe (1982) 42.
( Eds.), Stud. Surf. Sci. Catal., vol. 61, Elsevier, [94] J. Nováková, L. Kubelková, K. Habersberger, Z.
Amsterdam, 1991, p. 405. Dolejšek, J. Chem. Soc. Faraday Trans. I 80 (1984) 1457.
[64] S. Kolboe, in: A. Holmen, K.-J. Jens, S. Kolboe ( Eds.), [95] N.Y. Chen, W.J. Reagan, J. Catal. 59 (1979) 123.
Stud. Surf. Sci. Catal., vol. 61, Elsevier, Amsterdam, [96 ] C.D. Chang, Chem. Engng Sci. 35 (1980) 619.
1991, p. 413. [97] R.G. Anthony, Chem. Engng Sci. 36 (1981) 789.
[65] I.M. Dahl, S. Kolboe, J. Catal. 149 (1994) 458. [98] U. Sedran, A. Mahay, H.I. De Lasa, Chem. Engng Sci.
[66 ] I.M. Dahl, S. Kolboe, Catal. Lett. 20 (1993) 329. 45 (1990) 1161.
44 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

[99] R.G. Anthony, B.B. Singh, Chem. Engng Comm. 6 Innovation in Zeolite Materials Science, Elsevier,
(1980) 215. Amsterdam, 1987, p. 393.
[100] R. Mihail, S. Straja, G. Maria, G. Musca, G. Pop, Ind. [128] C.D. Chang, C.T.-W. Chu, R.F. Socha, J. Catal. 86
Engng Chem. Process Res. Div. 22 (1983) 532. (1984) 289.
[101] R. Mihail, S. Straja, G. Maria, G. Musca, G. Pop, Chem. [129] G. Chen, J. Liang, Q. Wang, G. Cai, S. Zhao, H. Li, in:
Engng Sci. 38 (1983) 1581. D.M. Bibby, C.D. Chang, R.F. Howe, S. Yurchak (Eds.),
[102] G.J. Hutchings, F. Gottschalk, R. Hunter, Ind. Engng Methane Conversion, Elsevier, Amsterdam, 1988, p. 201.
Chem. Res. 26 (1987) 637. [130] R.L. Espinoza, Appl. Catal. 26 (1986) 203.
[103] S.N. Tshabalala, A.M. Squires, AIChE J. 42 (1996) 2941. [131] D. Prinz, L. Riekert, Appl. Catal. 37 (1988) 139.
[104] D.B. Luk’yanov, Kinetics Catal. 30 (1989) 216. [132] U. Hammon, M. Kotter, L. Riekert, Appl. Catal. 37
[105] F.X. Cormerais, Y.-S. Chen, M. Kern, N.S. Gnep, G. (1988) 155.
Perot, M. Guisnet, J. Chem. Res. S (1981) 290. [133] W.J.H. Dehertog, G.F. Froment, Appl. Catal. 71
[106 ] T. Sodesawa, React. Kinetics Catal. Lett. 32 (1986) 251. (1991) 153.
[107] M. Krivánek, N.T. Dung, P. Jiru, Coll. Czech. Chem. [134] Y. Ono, E. Emai, T. Mori, Zeits. Phys. Chemie N.F. 115
Comm. 52 (1987) 1701. (1979) 99.
[108] L. Kubelková, J. Nováková, P. Jiru, in: P.A. Jacobs, N.I. [135] L. Jansen van Rensburg, R. Hunter, G.J. Hutchings,
Jaeger, P. Jiru, V.B. Kazansky, G. Schulz-Ekloff (Eds.), Appl. Catal. 42 (1988) 29.
Structure and Reactivity of Modified Zeolites, Elsevier, [136 ] C.D. Chang, W.H. Lang, R.L. Smith, J. Catal. 56
Amsterdam, 1984, p. 217. (1979) 169.
[109] B. Sulikowsky, A. Popielarz, Appl. Catal. 42 (1988) 195. [137] C.D. Chang, W.H. Lang, US Patent 4 025 576, 1977.
[110] N. Davidova, D. Shopov, N.I. Jaeger, G. Schulz-Ekloff, [138] W.O. Haag, R.M. Lago, P.G. Rodewald, J. Mol. Catal.
React. Kinetics Catal. Lett. 12 (1979) 229. 17 (1982) 161.
[111] S. Hočevar, G.V. Echevskii, B. Drzaj, K.J. Ione, React. [139] P.D. Caesar, R.A. Morrison, US Patent 4 083 888, 1978.
[140] L. Marosi, J. Stabenow, M. Schwarzmann, Deutsche
Kinetics Catal. Lett. 13 (1980) 425.
Offenlegungsschrift DE 2827 385, 1980.
[112] A.B. Schwartz, J. Ciric, in: P.V. Venuto, P.S. Landis
[141] S.A. Tabak, US Patent 4 482 772, 1984.
( Eds.), Advanced Catalysis, vol. 18, 1968, 259.
[142] M.M. Wu, W.W. Kaeding, J. Catal. 88 (1984) 478.
[113] C.S. Lee, G.E. Stead, EP Application 105 591 A1 (1983).
[143] B.P. Pebrine, US Patent 4 100 262, 1978.
[114] E.J. Rosinski, M.K. Rubin, M.M.-S. Wu, EP Application
[144] J. Heering, L. Riekert, L. Marosi, in: D.H.Olson, A. Bisio
112 071 A1, 1983.
( Eds.), Proceedings of the 6th International Zeolite
[115] B. Juguin, F. Raatz, C. Travers, G. Martino, C. Hamon,
Conference, Butterworths, Surrey, 1984, p. 528.
French Patent 2 633 924, 1988.
[145] K. Suzuki, Y. Kiyozumi, K. Matsuzaki, S. Shin, Appl.
[116 ] P. Dejaifve, A. Auroux, P.C. Gravelle, J.C. Vedrine, Z.
Catal. 42 (1988) 35.
Gabelica, E.G. Derouane, J. Catal. 70 (1981) 123.
[146 ] L. Juan, in B. Drzaj, S. Hočevar, S. Pejovnik ( Eds.),
[117] M. Sawa, M. Niwa, Y. Murakami, Chem. Lett. 8 Zeolites, Elsevier, Amsterdam, 1985, p. 611.
(1987) 1637. [147] P.G. Rodewald, US Patent 4 066 714, 1978.
[118] M. Sawa, M. Niwa, Y. Murakami, Appl. Catal. 53 [148] C.D. Chang, C.T.-W. Chu, European Patent Application
(1989) 169. 123 449 A1, 1984.
[119] M. Sawa, K. Kato, K. Hirota, Y. Murakami, Appl. [149] K. Suzuki, Y. Kiyozumi, K. Matsuzaki, S. Shin, Appl.
Catal. 64 (1990) 297. Catal. 35 (1987) 401.
[120] G. Martino, B. Juguin, Deutsche Offenlegungsschrift DE [150] H. Shoji, Japanese Patent Application 59 219 134, 1984.
3604636 A1, 1986. [151] S. Nowak, Freiberger Forschungshefte., Reihe A 763
[121] C. Hamon, J. Bandiers, M. Senes, US Patent 4 447 669, (1987) 116.
1984. [152] T. Inui, T. Suzuki, M. Inone, Y. Murakawi. Y.
[122] N.V. Kljueva, N.D. Tien, K.G. Ione, Acta Phys. Chem. Takagami, in: P.A. Jacobs, N.I. Jaeger, P. Jiru, V.B.
31 (1985) 525. Kazansky, G. Schulz-Ekloff ( Eds.), Structure and
[123] M. Sawa, M. Niwa, Y. Murakami, 62nd CATSJ Meeting Reactivity of Modified Zeolites, Elsevier, Amsterdam,
Abstracts 30 (1988) 436. 1984.
[124] O.V. Kikhtyanin, V.M. Mastikhin, K.G. Ione, Appl. [153] J.C. Oudejans, P.F. van den Oosterkamp, H. van
Catal. 42 (1988) 1. Bekkum, Appl. Catal. 3 (1982) 109.
[125] N.J. Tapp, N.B. Milestone, L.J. Wright, J. Chem. Soc. [154] R.A. Rajadhyaksha, J.R. Anderson, J. Catal. 63 (1980)
Chem. Comm. (1985) 1801. 510.
[126 ] K.J. Chao, L.J. Leu, in: H.G. Karge, J. Weitkamp (Eds.), [155] V.R. Choudhary, V.S. Nayak, Zeolites 5 (1985) 15.
Zeolites as Catalysts, Sorbents and Detergent Builders, [156 ] W.W. Kaeding, S.A. Butter, US Patent 3 911 041, 1975.
Elsevier, Amsterdam, 1989, p. 19. [157] W.W. Kaeding, S.A. Butter, J. Catal. 61 (1980) 155.
[127] N.J. Tapp, N.B. Milestone, D.M. Bibby, in: P.J. Grobet, [158] J.C. Vedrine, A. Auroux, P. Dejaifve, V. Ducarme, H.
W.J. Mortier, E.F. Vansant, G. Schulz-Ekloff ( Eds.), Hoser, S. Zhou, J. Catal. 73 (1982) 147.
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 45

[159] I. Balkrishnan, B.S. Rao, S.G. Hegde, A.N. Kotasthane, [182] T. Inui, H. Matsuda, O. Yamase, H. Nagata, K. Fukuda,
S.B. Kulkarni, P. Ratnaswamy, J. Mol. Catal. 17 T. Ukawa, A. Miyamoto, J. Catal. 98 (1986) 491.
(1982) 261. [183] V.N. Romannikov, L.S. Chumachenko, V.M. Mastikhin,
[160] G. Cai, G. Chen, Q. Wang, Q. Xin, X. Wang, X. Li, K.G. Ione, J. Catal. 94 (1985) 508.
J. Liang, in: B. Drzaj, S. Hočevar, S. Pejovnik ( Eds.), [184] T. Fleckenstein, K. Belendorff, F. Fetting, Chem. Ing.
Zeolites, Elsevier, Amsterdam, 1985, p. 319. Technol. 57 (1985) 800.
[161] J. Liang, G. Chen, S. Zhao, Q. Wang, H. Li, G. Cai, [185] T. Fleckenstein, K. Belendorff, F. Fetting, Ger. Chem.
Chem. Express 1 (1986) 729. Ing. 9 (1986) 346.
[162] S.A. Butter, US Patent 3 979 472, 1976. [186 ] A.A. Avidan, R.M. Gould, S.E. Kane, US Patent
[163] P.G. Rodewald, US Patent 4 145 315, 1979. 4 547 616, 1985.
[164] P.G. Rodewald, US Patent 4 100 219, 1978. [187] N. Daviduk, J.H. Haddad, US Patent 4 423 274, 1983.
[165] K.G. Ione, L.A. Vostrikova, A.V. Petrova, V.M. [188] A.A. Avidan, A.Y.-Y. Kam, European Patent
Mastikhin, in: P.A. Jacobs, N.I. Jaeger, P. Jiru, V.B. Application 120 619 A1, 1984.
Kazansky, G. Schulz-Ekloff ( Eds.), Structure and [189] R.F. Socha, C.D. Chang, R.M. Gould, S.E.Kane, A.A.
Reactivity of Modified Zeolites, Elsevier, Amsterdam, Avidan, Symp. Chem. Syng. Methanol, Div. Petr. Chem.
1984, p. 151. Inc., American Chemical Society, New York, 1986, p. 16.
[166 ] M.R. Klotz, US Patent 4 292 458, 1981. [190] C.H. Hsia, H. Owen, B.S. Wright, US Patent 4 506 106,
[167] W. Hölderich, H. Eichhorn, R. Lehnert, L. Marosi, W. 1985.
Mross, R. Reinke, W. Ruppel and H. Schlimper in D.H. [191] H. Owen, S.A. Tabak, B.S. Wright, US Patent
Olson and A. Bisio ( Eds.), Proceedings of the 6th 4 634 798, 1987.
International Zeolite Conference, Butterworths, Surrey [192] H. Owen, S.A. Tabak, B.S. Wright, US Patent
( UK ), 1984, p. 545. 4 628 135, 1986.
[168] W. Hölderich, W. Mross, M. Schwarzmann, Deutsche [193] K.R. Graziani, A.V. Sapre, US Patent 4 542 252, 1985.
Offenlegungsschrift DE 3136 984 A1 1983. [194] S.A. Tabak, US Patent 4 654 453, 1987.
[169] O. Forlani, Deutsche Offenlegungsschrift DE [195] L.B. Young, US Patent 4 433 189, 1984.
3540 286 A1, 1986. [196 ] W.W. Kaeding, European Patent Application
[170] C.T.-W. Chu, G.H. Kuehl, R.M. Lago, C.D. Chang, 229 952 A3, 1986.
J. Catal. 93 (1985) 451. [197] C.D. Chang, A.J. Silvestri, R.L. Smith, US Patent
[171] G. Coudurier, A. Auroux, J.C. Vedrine, R.D. Farlee, L. 3 928 483, 1975.
Abrams, R.D. Shannon, J. Catal. 108 (1987) 1. [198] R.M. Lago, US Patent 4 025 571, 1977.
[172] M.B. Sayed, A. Auroux, J.C. Vedrine, J. Catal. 116 [199] S.A. Butter, A.T. Jurewicz, W.W. Kaeding, US Patent
(1989) 1. 3 894 107, 1975.
[173] J.C. Jansen, E. Biron and H. van Bekkum in P.J. Grobet, [200] M.N. Harandi, H. Owen, US Patent 4 835 329, 1989.
W.J. Mortier, E.F. Vansant, G. Schulz-Ekloff ( Eds.), [201] K.R. Graziani, A.V. Sapre, US Patent 4 550 217, 1985.
Innovation in Zeolite Materials Science, Elsevier, [202] R.M. Dessau, G.T. Kerr, US Patent 4 642 407, 1987.
Amsterdam, 1987, p. 133. [203] E. Bowes, C.D. Chang, R.F. Socha, European Patent
[174] B. Andersen, C.T. O’Connor, M. Kojima: in P.A. Jacobs, Application 172 686 A1, 1985.
R.A. van Santen ( Eds.), Zeolites: Facts, Figures, Future, [204] C.T.-W. Chu, C.D. Chang, J.N. Miale, European Patent
Elsevier, Amsterdam, 1989, p. 1193. Application 114 498, 1983.
[175] S.M. Csicsery, Pure Appl. Chem. 58 (1986) 841. [205] C.T.-W. Chu, European Patent 252 742 B1, 1987.
[176 ] H. Baltes, H. Litterer, E.I. Leupold, F. Wunder, Deutsche [206 ] D. Timm, K. Wehner, H.J. Derdulla, H.Striegler, DDR
Offenlegungsschrift DE 3141 285 A1, 1983. Patent DD 257 740 A3, 1988.
[177] H. Baltes, H. Litterer, E.I. Leupold, F. Wunder, Deutsche [207] M. Weber, K. Becker, R. Thätner, H. Kieser, P. Birka,
Offenlegungsschrift DE 3141 283 A1, 1983. M. Prag, O. Rademacher, E. Rönsch, H. Scheler, K.
[178] U. Dettmeier, H. Baltes, H. Litterer, E.I. Leupold, W. Voigtberger, A. Göbler, H. Tschritter, D. Timm, H.
Herzog, F. Wunder, Chem. Ing. Technol. 54 (1982) 593. Striegler, DDR Patent DD 253 611 A1, 1988.
[179] T. Inui, O. Yamase, K. Fukuda, A. Itoh, J. Tarumoto, [208] C.D. Chang, J.N. Miale, US Patent 4 605 803, 1986.
M. Morinaga, T. Hagiwara, T. Takegami, in: Proceedings [209] S.G. Gagarin, Neftekhimiya 28 (1988) 208.
8th International Congress on Catalysis (Berlin), Verlag [210] E.G. Derouane, P. Dejaifve, Z. Gabelica, Faraday Disc.
Chemie, Weinheim, 1984, p. III-569. Chem. Soc. 82 (1981) 331.
[180] T. Inui, D. Medhanavyn, P. Praserthdam, K. Fukuda, [211] K. Foger, J.V. Sanders, D. Seddon, Zeolites 4 (1984) 337.
T. Ukawa, A. Sakamoto, A. Miyamoto, Appl. Catal. 18 [212] C.D. Chang, C.T.-W. Chu, P.D. Perkins, E.W. Valyocsik,
(1985) 311. US Patent 4 476 338, 1984.
[181] T. Inui, A. Miyamoto, H. Matsuda, H. Nagota, Y. [213] J.L. Casci, B.M. Lowe, T. Vincent, UK Patent GB
Makino, K. Fukuda, F. Okazumi, in: Y. Murakami, A. 2077 709, 1981.
Jijima, J.W. Ward (Eds.), New Developments in Zeolite [214] G.W. Dodwell, R.P. Denkewies, L.B. Sand, Zeolites 5
Science and Technology, Elsevier, Tokyo, 1986, p. 859. (1985) 153.
46 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

[215] C.D. Chang, W.H. Lang, A.J. Silvestri, US Patent [244] J.M.O. Lewis, W.H. Henstock, European Patent
4 062 905, 1977. 359 843 B1, 1988.
[216 ] R.G. Anthony, B.B. Singh, Hydrocarbon Processing [245] J.M.O. Lewis, W.H. Henstock, South Africa Patent
(1981) 85. 88/7237, 1988.
[217] S. Cartlidge, R. Patel, in: P.A. Jacobs, R.A. van Santen [246 ] J.M.O. Lewis, European Patent Application
( Eds.), Zeolites: Facts, Figures, Future, Elsevier, 359 842 A1, 1988.
Amsterdam, 1989, p. 1151. [247] A.J. Marchi, G.F. Froment, Appl. Catal. 71 (1991) 139.
[218] F.A. Wunder, E.I. Leupold, Angew. Chem. 92 (1980) [248] J.A. Rabo, Periodica Polytech., Chem. Engng 32
125. (1988) 211.
[219] U. Dettmeier, E.I. Leupold, H. Litterer, H. Baltes, W. [249] J. Liang, H. Li, S. Zhao, W. Guo, R. Wang, M. Ying,
Herzog, F.A. Wunder, Erdöl, Kohle, Erdgas, Petrochem. Appl. Catal. 64 (1990) 31.
36 (1983) 365. [250] S.W. Kaiser, European Patent Application EP 249 915,
[220] B.B. Singh, R.G. Anthony, Prepr. Can. Symp. Catal. 6 1987.
(1979) 113. [251] S.W. Kaiser, Norwegian Patent Application 872 505,
[221] B.B. Singh, F.N. Lin, R.G. Anthony, Chem. Engng 1987.
Comm. 4 (1980) 749. [252] T. Hibi, K. Takahashi, T. Okuhara, M. Misono, Y.
[222] G.V. Tsitsishvili, T.M. Ramishvili, M.K. Charkviani, in: Yoneda, Appl. Catal. 24 (1986) 69.
P.A. Jacobs, N.I. Jaeger, P. Jiru, V.B. Kazansky, G. [253] J.B. Moffat, in: D.M. Bibby, C.D. Chang, R.F. Howe,
Schulz-Ekloff ( Eds.), Structure and Reactivity of S. Yurchak ( Eds.), Methane Conversion, Elsevier,
Modified Zeolites, Elsevier, Amsterdam, 1984, p. 141. Amsterdam, 1988, p. 563.
[223] S. Ceckiewicz, Bull. Acad. Pol. Sci., Ser. Chim. 27 [254] H. Ehwald, W. Fiebig, H.-G. Jerschkewitz, G. Lischke,
(1979) 629. B. Parlitz, E. Schreier, G. Öhlmann, Appl. Catal. 34
[224] S. Ceckiewicz, React. Kinetics Catal. Lett. 16 (1981) 11. (1987) 13.
[225] D. Gubisch, F. Bandermann, Chem. Ing. Technol. 12 [255] H. Ehwald, W. Fiebig, H.-G. Jerschkewitz, G. Lischke,
(1989) 155. B. Parlitz, P. Reich, G. Öhlmann, Appl. Catal. 34
[226 ] E.N. Givens, C.J. Plank, E.J. Rosinski, US Patent (1987) 23.
4 079 095, 1978. [256 ] A.J. Cooper, F.A. Pesa, J.K. Currie, US Patent
[227] E.N. Givens, C.J. Plank, E.J. Rosinski, US Patent 4 757 044, 1988.
4 079 096, 1978. [257] M.J.G. Janssen, D.E.W. Vaughan, International Patent
[228] N.P. Forbus, M.M. Wu, US Patent 4 449 961, 1984. WO 93/24429, 1993.
[229] Y. Tsutsumi, Japanese Patent Application 58-74930, [258] A.J. Marchi, G.F. Froment, Appl. Catal. 94 (1993) 91.
1984. [259] S. Hočevar, J. Batista, V. Kaučič, J. Catal. 139 (1993)
[230] D. Timm, K. Becker, H.J. Derdulla, H. Striegler, M. 351.
Weber, DDR Patent DD 238 376 A1, 1986. [260] M. Jayamurthy, S. Vasudevan, Ber. Bunsenges. Phys.
[231] K.G. Ione, V.G. Stepanov, L.A. Vostrikova, SU Patent Chem. 99 (1995) 1521.
914 544, 1982. [261] M. Jayamurthy, S. Vasudevan, Catal. Lett. 36 (1996) 111.
[232] T. Inui, Y. Takegami, Hydrocarbon Processing (1982) [262] S.R. Blaszkowski, R.A. van Santen, J. Am. Chem. Soc.
117. 119 (1997) 5020.
[233] T. Inui, N. Morinaga, Y. Takegami, Appl. Catal. 8 [263] S.R. Blaszkowski, R.A. van Santen, J. Phys. Chem. B
(1983) 187. 101 (1997) 2292.
[234] T. Inui, T. Ishihara, Y. Takegami, J.C.S. Chem. Comm. [264] G. Mirth, J.A. Lercher, in: A. Holmen, K.-J. Jens, S.
(1981) 936. Kolboe (Eds.), Stud. Surf. Sci. Catal., vol. 61, Elsevier,
[235] G.J. Hutchings, T. Themistocleous, R.G. Copperthwaite, Amsterdam, 1991, p. 437.
Appl. Catal. 43 (1988) 133. [265] X. Bo-qing, L. Juan, C. Guo-quan, Z. Su-qin, W.
[236 ] H. Sakoh, M. Nitta, K. Aomura, Appl. Catal. 16 Rong-hui, Cuihua Xuebao 12 (1991) 273.
(1985) 249. [266 ] Y. Ying, D. Yingru, Ranliao Huaxue Xuebao 18
[237] M.S. Spencer, T.V. Whittam, J. Mol. Catal. 17 (1982) (1990) 247.
271. [267] K. Kawamura, Y. Kohno, K. Matsuzaki, T. Sano, H.
[238] A. Stewart, D.W. Johnson, M.D. Shannon, in: Takaya, Sekiyu Gakkaishi 34 (1991) 90.
P.J. Grobet, W.J. Mortier, E.F. Vansant, G. Schulz- [268] K. Kawamura, Japanese Patent Application 60-245435,
Ekloff (Eds.), Innovation in Zeolite Materials Science, 1987.
Elsevier, Amsterdam, 1987, p. 57. [269] K. Kawamura, Y. Kono, H. Okado, H. Takaya, US
[239] S.W. Kaiser, Arab. J. Sci. Engng 10 (1985) 361. Patent 4 926 006, 1990.
[240] S.W. Kaiser, US Patent 4 499 327, 1985. [270] T. Sano, Y. Kiyozumi, S. Shin, Sekiyu Gakkaishi 35
[241] S.W. Kaiser, US Patent 4 524 234, 1985. (1992) 429.
[242] S.W. Kaiser, US Patent 4 677 243, 1987. [271] M.G. Howden, J.J.C. Botha, M.S. Scurrell, Chem. Ind.
[243] S.W. Kaiser, Intern. Patent WO 86/04577, 1986. 46 (1992) 391.
M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48 47

[272] S.A. Tabak, S. Yurchak, Catal. Today 6 (1990) 307. ( Eds.), Stud. Surf. Sci. Catal., vol. 61, Elsevier,
[273] A.M. Al-Jarallah, U.A. El-Natafy, M.M. Abdillahi, Amsterdam, 1991, p. 421.
Appl. Catal. A: General 154 (1997) 117. [298] S. Nawaz, S.Kolboe, M. Stöcker, in: H.E. Curry-Hyde,
[274] A. Martin, B. Lücke, S. Nowak, H. Poethke, W. Wieker, R.F. Howe (Eds.), Natural Gas Conversion II, Elsevier,
B. Fahlke, U. Hahn, K. Anders, H. Günschel, H. Fürtig, Amsterdam, 1994, p. 393.
U. Hädicke, DDR Patent DD 270 296 A1, 1989. [299] A. Grønvold, K. Moljord, T. Dypvik, A. Holmen, in:
[275] H. Günschel, E. Thiele, S. Nowak, K. Anders, H. Fürtig, H.E. Curry-Hyde, R.F. Howe ( Eds.), Natural Gas
W. Wieker, H.-G. Vieweg, B. Lücke, H. Schulze, R. Conversion II, Elsevier, Amsterdam, 1994, p. 399.
Rehm, A. Martin, former DDR Patent DD 287 476 A5, [300] R. Wendelbo, D. Akporiaye, A. Andersen, I.M. Dahl,
now FRG Patent, 1991. H.B. Mostad, Appl. Catal. A: General 142 (1996) L197.
[276 ] M.N. Harandi, US Patent 5 146 032, 1992. [301] J.M.O. Lewis, W.H. Henstock, US Patent 4 973 792,
[277] M.N. Harandi, H. Owen, US Patent 4 981 491, 1991. 1990.
[278] G. Burgfels, K. Kochlöffl, J. Ladebeck, F. Schmidt, M. [302] N.N. Rostanin, L.D. Konoval’chikov, A.S. Shtefan, E.D.
Schneider, H.J. Wernicke, Deutsche Offenlegungsschrift Rostanina, B.K. Nefedor, Chimija i Technologija Topliv
DE 3838 710 A1, 1990. i Masel 3 (1992) 30.
[279] J.H. Beech, jr., F.P. Ragonese, US Patent 5 059 738, [303] X. Qinghua, C. Guoquan, W. Qingxia, W. Gougwei,
1991. Ranliao Huaxue Xuebao 22 (1994) 103.
[280] R.P.L. Absid, C.D. Chang, C.T.-W. Chu, D.J. Klocke, [304] Disclosed anonymously, Res. Disclosure Havant 328
US Patent 4 919 790, 1990. (1991) 634.
[281] S. Barri, D. Kidd, International Patent WO 93/02994, [305] E.G. Derouane, J.P. Gilson, J.B. Nagy, Zeolites 2
1993. (1982) 42.
[282] S. Barri, D. Kidd, European Patent Application [306 ] D.B. Luk’yanov, V.I. Timoshenko, M.G. Slin’ko, S.I.
485 145 A1, 1991. Ivanov, B.K. Nefedov, V.V. Vinogradov, Kinetics Catal.
[283] T. Inui, Sekiyu Gakkaishi 35 (1992) 33. 27 (1986) 1229.
[284] T. Inui, Am. Chem. Soc., Div. Petr. Chem. Prepr. 35 [307] D.M. Bibby, R.F. Howe, G.D. McLellan, Appl. Catal.
(1990) 124. A: General 93 (1992) 1.
[285] T. Inui, in: R.K. Grasselli, A.W. Sleight (Eds.), [308] M. Guisnet, P. Magnoux, Appl. Catal. 54 (1989) 1.
Structure–Activity and Selectivity Relationships in [309] L.D. Rollmann, J. Catal. 47 (1977) 113.
Heterogeneous Catalysis, Elsevier, Amsterdam, 1991, [310] L.D. Rollmann, D.E. Walsh, J. Catal. 56 (1979) 139.
p. 233. [311] H. Schulz, W. Böhringer, W. Baumgartner, Z. Siwei, in:
[286 ] T. Inui, Yuki Gosei Kagaku Kenkyusho Koenshu 4 Y. Murakami, A. Jijima, J.W. Ward (Eds.), New
(1990) 84. Developments in Zeolite Science and Technology,
[287] T. Inui, S. Phatanasri, H. Matsuda, Catal. Sci. Technol. Elsevier, Tokyo, 1986, p. 915.
1 (1991) 85. [312] A.A. Slinkin, A.V. Kucherov, D.A. Kondratyev, T.N.
[288] T. Inui, European Patent 418 142 B1, 1994. Bondarenko, A.M. Rubinstein, Kh. M. Minachev, in: Y.
[289] T. Inui, in: H. Chon, S.-K. Ihm, Y.S. Uh ( Eds.), Progress Murakami, A. Jijima, J.W. Ward (Eds.), New
in Zeolite and Microporous Materials, Elsevier, Developments in Zeolite Science and Technology,
Amsterdam, 1997, p. 1441. Elsevier, Tokyo, 1986, p. 819.
[290] G. Pop, G. Musca, D. Ivanescu, E. Pop, G. Maria, E. [313] P. Magnoux, P. Cartraud, S. Mignard, M. Guisnet,
Chirila, O. Munteau, Chem. Ind. 46 (1992) 443. J. Catal. 106 (1987) 242.
[291] L. Smith, A.K. Cheetham, L. Marchese, J.M. Thomas, [314] F.X. Cormerais, G. Penet, M. Guisnet, Zeolites 1
P.A. Wright, J. Chen, E. Gianotti, Catal. Lett. 41 (1981) 141.
(1996) 13. [315] K.G. Ione, G.V. Echevskii, G.N. Nosyreva, J. Catal. 85
[292] J.M. Thomas, Y. Xu, C.R.A. Catlow, J.W. Couves, (1984) 287.
Chem. Mater. 3 (1991) 667. [316 ] D.M. Bibby, N.B. Milestone, J.E. Patterson, L.P.
[293] M.J. van Niekerk, J.C.Q. Fletcher, C.T. O’Connor, Appl. Aldridge, J. Catal. 97 (1986) 493.
Catal. A: General 138 (1996) 135. [317] D.M. Bibby, G.D. McLellan, R.F. Howe, in: B. Delmon,
[294] P.T. Barger, US Patent 5 095 163, 1992. G.F. Froment ( Eds.), Catalyst Deactivation 1987,
[295] J. Chen, P.A. Wright, S. Natarajan, J.M. Thomas, in: Elsevier, Amsterdam, 1987, p. 651.
J. Weitkamp, H.G. Karge, H. Pfeifer, W. Hölderich [318] G.D. McLellan, R.F. Howe, L.M. Parker, D.M. Bibby,
( Eds.), Zeolites and Related Microporous Materials: J. Catal. 99 (1986) 486.
State of the Art 1994, Elsevier, Amsterdam, 1994, p. 1731. [319] B.A. Sexton, A.E. Hughes, D.M. Bibby, J. Catal. 109
[296 ] S.M. Yang, S.I. Wang, C.S. Huang, in: A. Holmen, (1988) 126.
K.-J. Jens, S. Kolboe ( Eds.), Stud. Surf. Sci. Catal., [320] T. Behrsing, H. Jaeger, J.V. Sanders, Appl. Catal. 54
vol. 61, Elsevier, Amsterdam, 1991, p. 429. (1989) 289.
[297] S. Nawaz, S. Kolboe, S. Kvisle, K.-P. Lillerud, M. [321] G.F. Froment, J. De Meyer, E.G. Derouane, J. Catal.
Stöcker, H. Øren, in: A. Holmen, K.-J. Jens, S. Kolboe 124 (1990) 391.
48 M. Stöcker / Microporous and Mesoporous Materials 29 (1999) 3–48

[322] G. Chen, J. Liang, Q. Wang, G. Cai, S. Zhao, M. Ying, [336 ] D. Chen, H.P. Rebo, K. Moljord, A. Holmen, in: C.H.
in: Y. Murakami, A. Jijima, J.W. Ward ( Eds.), New Bartholomew, G.A. Fuentes ( Eds.), Catalyst
Developments in Zeolite Science and Technology, Deactivation 1997, Elsevier, Amsterdam, 1997, p. 159.
Elsevier, Tokyo, 1986, p. 907. [337] D. Chen, K. Moljord, H.P. Rebo, A. Holmen (in press).
[323] B.E. Langner, Appl. Catal. 2 (1982) 289. [338] N.Y. Chen, W.E. Garwood, Catal. Rev., Sci. Engng 28
[324] G.V. Echevskii, N.G. Kalinina, V.F. Anufrienko, V.A. (1986) 185.
Poluboyarow, React. Kinetics Catal. Lett. 33 (1987) 305. [339] N.Y. Chen, W.O. Haag, Chem. Ind. (NY ) 31 (1988) 695.
[325] G. Pop, G. Musca, E. Chirila, R. Boeru, G. Niculae, N. [340] H.R. Grimmer, N. Thiagarajan, E. Nitschke, in: D.M.
Natu, G. Ignatescu, S. Straja, Chem. Ing. Sci. 44 Bibby, C.D. Chang, R.F. Howe, S. Yurchak ( Eds.),
(1989) 49. Methane Conversion, Elsevier, Amsterdam, 1988, p. 273.
[326 ] P.H. Nelson, D.M. Bibby, in: C.H. Bartholomew, J.B. [341] S. Nowak, H. Günschel, A. Martin, K. Anders, B. Lücke,
Butt ( Eds.), Catalyst Deactivation 1991, Elsevier, in: M.J. Phillips, M. Ternan (Eds.), Catalysis: Theory to
Amsterdam, 1991, p. 407. Practice, Proceedings 9th International Congress on
[327] F. Bauer, H. Ernst, E. Geidel, R. Hanisch, E. Onderka, Catalysis, The Chemical Institute of Canada, Ottawa,
R. Schödel, Erdöl, Erdgas, Kohle 112 (1996) 516. 1988, Vol. 4, p. 1735.
[328] A.T. Aguayo, P.L. Benito, A.G. Gayubo, M. Olazar, [342] A. Martin, S. Nowak, B. Lücke, H. Günschel, Appl.
J. Bilbao, in: B. Delmon, G.F. Froment ( Eds.), Catalyst Catal. 50 (1989) 149.
Deactivation 1994, Elsevier, Amsterdam, 1994, p. 567. [343] A. Martin, S. Nowak, B. Lücke, W. Wieker, B. Fahlke,
[329] P.L. Benito, A.G. Gayubo, A.T. Aguayo, M. Olazar, J. Appl. Catal. 57 (1990) 203.
Bilbao, Ind. Engng Chem. Res. 35 (1996) 3991. [344] J. Weitkamp, Stud. Surf. Sci. Catal. 65 (1991) 21.
[330] J.M. Ortega, A.G. Gayubo, A.T. Aguayo, P.L. Benito, [345] R.F. Socha, C.D. Chang, R.M. Gould, S.E. Kane, A.A.
J. Bilbao, Ind. Engng Chem. Res. 36 (1997) 60. Avidan, Am. Chem. Soc. Symp. Ser. 328 (1987) 34.
[331] S. Ceckiewicz, J. Chem. Soc. Faraday Trans. I 80 [346 ] B.V. Vora, T.L. Marker, 2nd International Petroleum
(1984) 2989. Conference and Exhibition, Petrotech 97, New Dehli,
[332] E.G. Derouane, in: B. Imelik, C. Naccache, G. India, 1997.
Coudurier, Y. Ben Taarit, J.C. Vedrine (Eds.), Catalysis [347] C.N. Eng, H.R. Nilsen, V. May, Asian Olefins and
by Acids and Bases, Elsevier, Amsterdam, 1985. Derivatives Conference, Singapore, June 18–19, 1997.
[333] J.W. Beeckman, G.F. Froment, Ind. Engng Chem. [348] European Chemical News, 19.08.1996, p. 26.
Fundament. 18 (1979) 245. [349] T. Inui, M. Kang, Appl. Catal. A: General 164 (1997)
[334] J.W. Beeckman, G.F. Froment, Chem. Engng Sci. 35 211.
(1980) 805. [350] P.W. Goguen, T. Xu, D.H. Barich, T.W. Skloss, W.
[335] D. Chen, H.P. Rebo, K. Moljord, A. Holmen, Ind. Engng Song, Z. Wang, J.B. Nicholas, J.F. Haw, J. Am. Chem.
Chem. Res. 36 (1997) 3473. Soc. 120 (1998) 2650.

You might also like