You are on page 1of 27

Applied Catalysis, 54 (1989) 1-27

Elsevier Science Publishers B.V., Amsterdam - Printed in The Netherlands

Review

Coking and Deactivation of Zeolites

Influence of the Pore Structure

M. GUISNET* and P. MAGNOUX


URA CNRS 00350, Catalyse en Chimie Organique, UniuersitP de Poitiers, 40, avenue du
Recteur Pineau, 86022 Poitiers Ceder {France)
(Received 17 February 1989)

CONTENTS

Abstract ..................................................................................................... 2
1. Introduction .............................................................................................. 2
1.1 Methods for coke characterization ................................................... 3
1.2 Objective ............................................................................................. 5
2. Pore structure. Coking and deactivation rates ...................................... 6
2.1 n-Heptane cracking: rate and selectivity ......................................... 6
2.2 Deactivation and coking rates .......................................................... 6
2.3 Pore structure, coking and aging rates ............................................ 8
2.3.1 Pore structure and coking rate ................................................ 8
2.3.2 Pore structure and deactivating effect of coke ....................... 9
2.4 Conclusion .......................................................................................... 9
3. Coke composition. Location, mode of formation .................................. 10
3.1 Characterization techniques ............................................................. 10
3.2 Coke composition ............................................................................... 11
3.2.1 Composition of the soluble coke .............................................. 11
3.2.2 Composition of the non-soluble coke ...................................... 14
3.3 Location of coke ................................................................................ .14
3.4 Mode of coke formation .................................................................... 15
4. Mode of deactivation ............................................................................... 17
4.1 Deactivating effect of the coke molecules ........................................ 17
4.2 Limitation or blockage of the access to the pores ........................... 19
4.3 Modes of deactivation ....................................................................... .21
4.4 Conclusion .......................................................................................... 22
5. Conclusion ................................................................................................ 24
References ................................................................................................. 25

0166-9834,‘89/$03.50 0 1989 Elsevier Science Publishers B.V.


2

ABSTRACT

The modes of coking and of deactivation of zeolites during n-heptane cracking at 723 K were
established on the basis of (i) the composition of the carbonaceous compounds responsible for
deactivation (coke), (ii) the deactivating effect of the coke molecules and (iii) the reduction by
coke of the volume accessible to nitrogen and to n-hexane (kinetic diameter similar to n-heptane).
The zeolites [USHY, H Mordenite (HMOR), HZSMS and H Erionite (HERI) ] were chosen to
determine the effect of different parameters of the pore structure: (i) pore size, (ii) existence
(USHY, HERI) or non-existence of cavities (HMOR, HZSM5), (iii) the possibility for the reac-
tant to diffuse unidirectionally (HMOR) or tridirectionally. The retention of coke molecules is
due to trapping in the cavities (or at channel intersections). Their size is intermediate between
that of the apertures and that of the cavities (or channel intersections). The coking rate is all the
faster when the space available near the acid sites is large and when the coke precursors desorb
slowly. On all the zeolites, coke formation occurs through oligomerization of the olefinic cracking
products followed by cyclization of the oligomers, transformation through hydrogen transfer into
monoaromatics, alkylation of these monoaromatics, then cyclization and hydrogen transfer to
give bi-aromatics, t&aromatics, etc. There is no site poisoning by coke; deactivation occurs through
the three following modes: (i) limitation of the access of n-heptane to the active sites, (ii) blockage
of the access to the sites of the cavities (or channel intersections) in which the coke molecules are
situated and (iii) blockage of the access to the sites of the pores in which there are no coke
molecules.

1. INTRODUCTION

All the reactions of organic compounds on solid catalysts are accompanied


by the formation of heavy by-products which form a deposit on the surface and
provoke deactivation. The deactivation rate depends obviously on the relative
rate of formation of these by-products, a rate which can be very different from
one catalytic system (catalyst + reaction) to another. Thus, in reforming, only
one atom of carbon out of 200 000 activated by the catalyst is transformed into
coke [ 11, while in the cracking of heavy petroleum cuts this can be more than
one out of 20 [ 2 1, The deactivation rate depends also on the mode of deacti-
vation: the deactivating effect of coke is obviously more pronounced when
deactivation is due to pore blockage than when it is due to site coverage [ 3,4].
Whatever the catalytic system it is important to limit the rate of deactivation.
To obtain this result it is essential to understand what is (are) the mode(s) of
coke formation as well as the mode (s) of deactivation.
Owing to their great activity, to their particular selectivity (shape selectiv-
ity) and to their high thermal stability, acid zeolites are more and more used
in industrial processes. This is the case for large pore-size zeolites such as Y
zeolite (cracking, hydrocracking, etc.) as well as for zeolites with smaller pore
sizes, e.g. ZSM5 (MTG, dewaxing, xylene isomerization, etc.) or Erionite
(selectoforming ).
In most of the industrial processes, the formation of heavy by-products, which
remain in the pores or form a deposit on the outer surface, is mainly responsible
for zeolite deactivation. Rollman and Walsh [5,6] are the first who have shown
that coke formation is a “shape-selective reaction”: the coking tendency is an
3

intrinsic property of the zeolite pore structure. Thus the smaller the pore size
the lower the coke yield in the cracking of a hydrocarbon mixture [ 51. While
this effect of the pore structure on the coking rate has been confirmed, it is
sometimes difficult to discriminate between this effect and that of the acidity:
for instance, the low coking rate found on HZSM5 which was first related only
to its pore structure is now attributed to a large part to the generally low den-
sity of its acid sites [ 71,
The pore structure has also a marked influence on the deactivating effect of
coke. Zeolites having a perfectly defined pore structure are ideal solids for
studying pore blockage and in particular for establishing relationships between
the characteristics of the pore structure and the deactivating effect of coke
molecules [ 81. A significant point is that the compounds responsible for zeolite
deactivation (coke) can be situated inside the micropores and/or on the outer
surface. The deactivating effect of coke depends obviously on this situat’ion.
Various techniques, including adsorption [ 7-201, X-ray photoelectron spec-
troscopy [ 211, electron microscopy [ 221 and X-ray diffraction [ 131, were used
to locate the coke inside the micropores or on the outer surface. Through ad-
sorption measurements with adsorbate molecules of different sizes it is even
possible to specify in which pores of zeolites with a double pore system, such
as offretite [ 141, the coke is located.
The major limit to our knowledge of the mode (s) of coking and of deacti-
vation is that of establishing the composition of coke. However, it is not easy
to determine the composition of coke because coke is a very complex mixture,
difficult to separate from the zeolite. This is why, generally, in situ methods of
characterizaton are used. However, these methods, while they give information
about the nature of coke (chemical identity), do not allow the distribution of
the coke components to be obtained. Before reviewing briefly the techniques
used for the characterization of coke in zeolites it is necessary to define what
we mean by the word “coke”. This term is generally used to designate the heavy
polynuclear aromatics responsible for the deactivation of most catalysts. The
by-products responsible for zeolite deactivation not being always polyaro-
matic, it is necessary either not to use the word coke or to give it a more general
definition. As the use of this word in works on deactivation is well established,
the second proposition seems to us to be the better. We shall use then in this
work the following definition of coke: carbonaceous compounds (polyaromatic
or non polyaromatic ) formed during a reaction and which are responsible for
deactivation [ 231.

1.1 Methods for coke characterization

Until recently the characteriiation of coke was limited often to the mea-
surement of its atomic hydrogen-to-carbon ratio. The information although
limit.ed is already significant. Thus it has been shown that the heavy by-prod-
4

ucts responsible for zeolite deactivation (coke) are not always polyaromatic
[24-281. However, the hydrogen-to-carbon value is often ambiguous in the
case of zeolitic catalysts. Indeed internal and external cokes, completely dif-
fering in compositions can coexist on the same zeolite) their relative signifi-
cance depending on the zeolite characteristics (pore structure, acidity) and on
operating conditions (temperature, reaction time).
Various spectroscopic techniques such as IR [ 24,27,29-391, generally Four-
ier transform (FT) IR, UV-VIS [ 401, lH cross polarization magic angle spin-
ning (CP/MAS) 13CNMR spectroscopy [ 41-46 1, electron paramagnetic res-
onance (EPR) [ 9,10,47-491 have been used for characterizing the chemical
nature of coke.
IR studies give information concerning the chemical identity of the coke
components (e.g. olefinic, saturated or aromatic), the amount of coke (e.g.
through the intensity of the so-called coke band around 1585 cm-l) and also
the changes in the catalyst characteristics (e.g. concentration of the Brcansted
OH sites). While most of the studies have been carried out under static con-
ditions, Karge et al. [ 29-32 ] have recently developed an on-stream technique.
UV-VIS Spectroscopy is also suitable for specifying the chemical identity of
the coke components. Thus, with this technique it was found that polyenylic
compounds were formed during ethylene oligomerization over dealuminated
H Mordenites [ 401.
CP/MAS 13C NMR spectroscopy is an efficient tool for investigating the
nature of carbonaceous deposits on zeolites. Derouane et al. [41] used this
technique to characterize the carbonaceous compounds formed during the
transformation of methanol and of ethylene on HZSM5 and on H mordenite
catalysts. Large differences were found between the two zeolites: mainly ali-
phatics and alkyl benzenics were found on HZSM5 instead of polyaromatics
on HMOR.
EPR spectroscopic studies of the radicals accompanying the formation of
coke on zeolites allow one to estimate the amount of coke and to obtain infor-
mation concerning its nature. EPR measurements can be made under static
and under on-stream conditions. Radicals observed during the formation of
coke through the reaction of ethylene or of propene on HMOR were found to
be typical of the coke species formed. Thus olefinic and allylic radicals accom-
panied the formation of low-temperature coke which was mainly composed of
olefinic oligomers or polymers while polyaromatic radicals accompanied the
formation of high-temperature coke which was mainly aromatic [ 491.
Various techniques have been used to recover the carbonaceous compounds
trapped in the zeolite pores. The direct treatment of the coked zeolite by an
organic solvent [28, 501 or under an inert gas flow at high temperature [51]
allows the recovery of a small part of the carbonaceous compounds. The zeolite
structure can be destroyed by ball-milling for a long time [ 52 ] and soluble coke
components extracted with an organic solvent. The internal coke compounds
5

can be liberated by solubilizing the zeolite in acid solutions [11,24,28,53,54].


The soluble coke components were analyzed by classical methods such as gas
chromatography (GC ) , high-performance liquid chromatography (HPLC ) , ‘H
NMR, IR, mass spectrometry (MS) etc. In this case the distribution of these
components (and not only their nature ) was obtained.

1.2 Objective

The aim of this paper is to show how the coking and the deactivation rates
as well as the composition of coke depend on the pore structure of zeolites. For
this, four protonic zeolites were chosen: two large-pore size zeolites, USHY
and H Mordenite (HMOR), one with intermediate-pore size, HZSM5, and one
with small pore apertures, H Erionite (HERI). With these zeolites it was pos-
sible to determine the effect of different parameters of the pore structure: (i)
pore size (ii) existence (USHY, HERI) or non-existence (HMOR, HZSM5)
of cavities and (iii) the possibility of the reactant and/or the products diffus-
ingunidirectionally (e.g. n-heptane in HMOR) or tridirectionally (e.g. n-hep-
tane in USHY) . Naturally the zeolites had not the same acidity: the density
and the strength of their acid sites were different (Table 1) . Therefore certain
differences in the coking and deactivation phenomena could probably be due
to the differences in acidity. However, the percentages of the protonic ex-
change of the zeolites were chosen so as to have similar initial activities for n-
heptane cracking at 723 K (the reaction used here to study coking and
deactivation).

TABLE 1

Unit cell formula and acidity of the zeolites.

Zeolites Formula nA,” nA,= AQ


USHY Na~.~H,*.,Al,,.sSi,~~.~O~~ 18.5 3.4 120

HMOR Na,.,H,.,Al,.,Si,*.30,, 10.8 7.5 135

HZSMS Nao.oolHz.lAl,.,Sig,.~O~~ 2.2 1.6 145

HER1 Ko.sN~.2H,.,A~.2S1,7.8
. 0 6s 12.4 8.0 140

“Total number of acid sites (10” g-l)


‘Number of acid sites on which the.heat of NH, adsorption is greater than 100 kJ mol-’ (10”
g-‘)
“Average heat of NH, adsorption for the A2 sites (kJ mol- ’ ).
6

2. PORE STRUCTURE. COKING AND DEACTIVATION RATES

2.1 n-Heptane cracking: rate and selectivity

While the initial cracking activities of the zeolites (after two minutes reac-
tion) are similar (55-65. lop3 mol h-’ g-‘) the selectivities are quite different
[ 7,&X33]. While C, and C, are always the major products (70-90% ) the ratios
C,-to-&, branched-to-linear products and olefins-to-alkanes (o/s) are quite
different (Table 2). With large pore zeolites, C4-to-C3 is close to 1, isoC,-to-
nC, relatively high and o/s well below 1. These values are those which can be
expected from a cracking reaction occurring through carbenium ions:
iC;- +C;
iC,+ + CT
nC&+nC,’ --+isoC++
nC& + C;
nC,’ + C;

and in which the olefinic products undergo rapid secondary transformation.


The selectivities of HER1 and HZSM5 (Table 2) can also be explained by this
mechanism provided it is admitted that:
(1) Branched molecules such as those of isoC, either remain blocked in the
cages with narrow openings (HER1 ) or desorb with difficulty (HZSM5 ) .
(2) The secondary reactions of the olefinic products are slower than those
occurring in large pore zeolites (higher o/s).
These selectivities could also be due to high field gradients arising from the
constraining porosity of these two zeolites and in the case of HZSM5 from its
low charge density also [ 59,601.

2.2 Deactivation ati coking rates

The deactivation rates of the zeolites are very different (Fig, 1). HZSM5
deactivates very slowly whereas, on the contrary, HMOR and above all HER1

TABLE 2

n-Heptane cracking

Initial cracking (A,) and coking activities (Ah) and initial selectivities of the various zeolites.

Zeolites A, GIG iC,/nC, o/s’ Ak A,IAc


10e3 mol h-’ g-’ 10e3 mol h-’ g-’

USHY 65 1 2.8 0.2 8.7 0.13


HMOR 56 1 1.7 0.3 8.0 0.14
HZSM5 55 0.7 0.55 0.7 0.05 0.001
HER1 60 0.35 0.05 0.55 6.3 0.11

“Ratio of olefins-to-alkanes
Ac 110-3ml,h‘1~1~

.
I
0.5 1 6 t(h)

Fig. 1. n-Heptane cracking on USHY ( l ), HMOR (k), HZSMS (*) and HER1 ( A ). Change
in the activity (A,, 10e3 mol h-’ g-‘) as a function of time-on-stream (t, h).

,
2 4 6 ttil,
Fig. 2. Percentage of coke (% C) deposited on USHY (0 ), HMOR (*), HZSM5 ($) and HERI
(A ) versus time-on-stream (t, h).

5 10 AC (XI
Fig. 3. n-Heptane cracking on USHY (0 ), HMOR (*), HZSMS (* ic, and HERI ( A ) Change
in the residual activity AR versus dC, the difference between the coke percentage and the coke
percentage after 2 minutes’ reaction.
rapidly lose nearly all their activity. USHY has an intermediate behaviour. The
low deactivation rate of HZSM5 is mainly due to the low coking rate. Indeed
on this zeolite the coke formation is initially about 100 times slower th.an on
the other zeolites (Fig. 2 ). For the other zeolites the initial coking rates are
quite similar. The differences observed in aging rates are therefore due to dif-
ferences in the deactivating effect of coke: the‘ “toxicity” of coke expressed in
grams of zeal&e deactivated by gram of coke is about 40 times greater on HERI,
15 times on HMOR, 3 times on USHY than on HZSM5 (Fig. 3).
It has to be noted that with all the zeolites the rate of coking decreases much
.more rapidly than does the rate of n-heptane cracking: thus on WHY the
coking-to-cracking rate ratio changes from 0.13 after 2 min to 0.01 after 6 h.

2.3 Pore structure, cokirzg and aging rates

2.3.2 Pore structure and coking rate


Initially, the coking rates as well as the ratios of the coking-to-cracking rates
are very similar on the large pore zeolites (USHY and HMOR) and on HERI.
This means that either their pore structures identically affect the coking rates
or that the effect of the pore structure is compensated for by that of the acidity.
The first hypothesis is not to be excluded. Thus steric constraints during the
formation of coke molecules are more pronounced in the relatively narrow er-
ionite cages than in the large Y supercages; but the limitation in the coking
rate caused by these constraints is probably compensated for by the “trap”
effect of the erionite cages on various molecules. The same explanation is valid
for HMOR: the coke molecules can only grow along the length of the channels
hence the coking rate is lower than on USHY; however the diffusion of coke
precursors is slower in HMOR than in USHY which has a positive effect on
the coking rate. One can also envisage a compensation for the effect of porosity
by an effect of acidity since the average strength and the density of the acid
sites differ from one zeolite to the other.
It is moreover probable that the lower coking rate found with HZSM5 is due
for a large part to the lower density of its acid sites (Table 1). Indeed on a
dealuminated HY zeolite having a density of strong acid sites close to that of
HZSM5, the coking rate is equal to 0.04*10-3 mol h-’ g-l and the ratio of
coking-to-cracking rates is 0.002, that is to say close to that found on HZSM5
[X4]. However, steric constraints exerted by the pore network of HZSM5 on
the formation of the bulky intermediates of coking can also be responsible for
this low coking rate [ 51.
It will always be difficult to evaluate with precision the effect of the pore
structure on the coking rate since it is impossible to obtain the same acidity
for zeolites of different pore structure. However, it is obvious that the coking
rate is all the greater when:
9

(1) The space available for its formation (e.g. in the cavities or at the channel
intersections) is greater.
(2) The intermediates to coke formation diffuse more slowly into the gas phase.

2.3.2 Pore structure and deactivating effect of coke


The deactivating effect of coke depends very much on the zeolite (Fig. 3).
This can be for a large part attributed to the differences in the pore structure.
Thus the very pronounced deactivating effect found with HMOR can be re-
lated to its monodimensional pore structure. The presence of one coke mole-
cule in a channel makes all the acid sites of this channel inaccessible to the
reactant from the end at which the molecule is located [ 8,561. The explanation
for the still more pronounced deactivating effect found on HERI, a zeolite with
a tridimensional pore network, is obviously different. Here the coke molecules
are formed preferentially in the cages close to the outer surface and block the
access to the inner cages. Indeed n-heptane cracking produces branched mol-
ecules which remain trapped in the erionite cages, thus limiting the diffusion
to the inner cages of reactant molecules, These branched molecules lead to
polyaromatic molecules which block the diffusion of the reactant molecules
[571.
Coke also has a more pronounced deactivating effect on USHY than on
HZSM5. Both these zeolites having tridimensional pore structures and, unlike
HERI, not having cavities (or channel intersections) capable of significantly
trapping the molecules, it is difficult to give an explanation based merely on
the differences in pore structure. The eventual effect of the pore structure will
be rediscussed later in the light of the coke composition. The more pronounced
deactivating effect found on USHY can be explained to a certain extent by the
heterogeneous distribution according to the strength of its acid sites [ 611, Coke
is formed preferentially on the strongest acid sites and causes their deactiva-
tion. Since these sites are the most active, the initial deactivating effect of coke
will be more pronounced than if all the active sites were of the same strength
as is the case on HZSM5 [ 62-641. The deactivating effect of coke will decrease
when the coke content decreases. This is what is observed on USHY (Fig. 3 ).

2.4 Conclusion

The four zeolites adjusted to have similar initial activities for n-heptane
cracking present different selectivities. However in every case, the olefin-to-
alkane ratio is below 1 indicating that coke formation occurs through second-
ary transformation of olefinic products. The coking rate and above all the deac-
tivating effect of coke are determined by the pore structure. In particular zeo-
lites with monodimensional pore structure or having large cavities with small
apertures (trap cavities) are highly sensitive to deactivation. However acidity
also plays a significant role and it is sometimes difficult to discriminate be-
tween its effect and that of the pore structure.

3. COKE COMPOSITION. LOCATION, MODE OF FORMATION

3.1 Characterization techniques

We have developed a technique for characterizing the compounds responsi-


ble for the zeolite deactivation (“coke”). The technique (Fig. 4) consisted of
treating the coked zeolites at room temperat.ure with a solution of hydrofluoric
acid at 40% in order to dissolve the zeolite and to liberate the inner “coke”. We
verified that this treatment does not cause any transformation of the coke
[ 111. The soluble components of “coke” extracted by methylene chloride were
analyzed by classical techniques such as GC, HYLC, HNMR and MS, while
the non-soluble coke was characterized chemically (atomic hydrogen-to-car-
bon ratio) and physically (electron microscopy, electron energy loss
spectroscopy).

Coked zeolite

HF Solution
40 %

I
Coke

Nature

Fig. 4. Coke characterization.


11

3.2 Coke composition

The composition of coke on the various zeolites will be compared for iden-
tical coke contents rather than for identical times on stream. These two pa-
rameters are obviously interconnected but in a different way from one zeolite
to the other.
On all the zeolites the hydrogen-to-carbon ratio decreases with increasing
time on-stream, i.e. when the coke content increases. For low coke contents
coke is non-aromatic (the hydrogen-to-carbon ratio is > 1) , except on HMOR.
At high contents, it is polyaromatic. Coke is always more aromatic for large
pore zeolites than for small or intermediate size pores (Fig. 5).
The solubility of coke in methylene chloride (R) depends also on the zeolite
(Fig. 6). However, except with HMOR the insoluble coke is not a primary
product of the coking reaction. At low coke content, R is equal to 100% for
USHY and for HZSM5. For HER1 the low R value is not due to the presence
of insoluble coke but to the elimination of highly volatile compounds (e.g. iso-
butene or toluene) during dissolving of the zeolite in hydrofluoric acid. With
USHY, HZSM5 and HERI, insoluble coke is observed for higher coke content,
its formation occurring for different coke contents according to the zeolite
(about 2% on USHY and on HER1 ,4% on HZSM5).

3.2.1 Composition of the soluble coke

The soluble coke was analyzed by GC, HPLC, HNMR, IR, MS and GC-MS.
The composition depends on the zeolite and for a given zeolite it changes with
the coke content. Thus with USHY, HZSM5 and HER1 the number of aro-

H/C

2-

\
1-L\ _ .

1 I I
5 10 l5 zc

Fig. 5. Atomic hydrogen-to-carbon ratio (H/C) versus the coke percentage (%C). USHY (@ ),
HMOR (*,),HZSMS ($)andHERI (A).
Fig. 6. Yield of coke recovered in methylene chloride (R) as a function of the coke content (% C 1
on USHY (a), HMOR (Jr), HZSM5 (*) andHER (A).

“AR
I

5 10 15 ‘xc

Fig. 7. Analysis of the soluble coke by H NMR: Change in the percentages of aromatic protons
WAR) versus the coke percentage (% C). USHY ( l ), HMOR (*), HZSM5 (*e) and HER1
(A).

volatile

USHY HNOR HZSNS HERI

Fig. 8. Analysis of the soluble coke by mass spectroscopy. Number of carbon atoms (n,) per mol-
ecule of the main components for low content, open bars (1-2 wt.-%); and for high content,
hatchedbars (5-9 wt.-%).
13

TABLE 3

Main components of the soluble coke for low and high coke contents; size and boiling point

Zeolites Coke content

USHY 2 wt.-% 9 wt.-%

(8.5x9.5 A) (673-723 K) (8.5X 12 A) (793 K)

/\ -,
HMOR 2 wt.-% 4.5 wt.-%

w \
(6.5x8.5 A) (613 K) (6x 12 A, (573 K-623 K)

HZSM5 1 wt.-% 7 wt.-%

EF
R

0
0
0

(10 A, (473 K-523 K) (8.5X8.5& (673 K)

HER1 0.5 wt.-% 6 wt.-%

CH3
-3 0 &p
(6.5 A) (384 K) (6.5x12.5A) (723 K)

matic rings in the components estimated by HPLC increases with the coke
content; on the contrary on HMOR the soluble coke, poorly aromatic at low
coke contents, becomes essentially non-aromatic at high contents. HNMR these
changes in aromaticity with coke content (Fig, 7). The number of carbon at-
oms of the main components also depends on the zeolite and on the coke con-
tent (Fig. 8).
The main components of the soluble coke found at low and at high coke
contents as well as their size and the approximate value of their boiling point
under normal pressure are indicated in Table 3.
14

“k

Fig. 9. Change of the number of soluble coke molecules ( nk, 10”’ g-‘) as a function of the coke
percentage (% C).USHY (O), HMOR (*),HZSM5 (*) andHER (A).

The number of soluble coke molecules per gram of zeolite was estimated for
different coke contents on every zeolite. Fig. 9 shows that this number passes
through a maximum for a given coke content. The soluble coke molecules are
therefore intermediates in the formation of insoluble coke, which was what
allowed us to suppose:
(i ) that insoluble coke is not formed at low coke content and
(ii) that the aromaticity of the soluble coke increases with the coke content
(indeed insoluble coke is composed of highly polyaromatic molecules ).

3.2.2 Composition of the non-soluble coke


The insoluble coke is in the form of black particles. Its insolubility in meth-
ylene chloride and its colour indicate that its components are polyaromatic.
This is confirmed by the atomic hydrogen-to-carbon ratio: 0.4-0.5 on USHY,
HZSM5 and HERI and 0.6 on HMOR.
By using transmission electron microscopy (TEM) and electron energy loss
spectroscopy (EELS) some information about the location of the insoluble
coke molecules and about their nature can be obtained. The study was carried
out only on USHY and on HZSM5 [x2]. With HZSM5 insoluble coke forms
an external envelope around the zeolite crystal which remains as an empty
mould after zeolite solubilization. EELS shows that its structure is similar to
that of coronene (pregraphitic). For USHY part of the coke is under the form
of 1 nm large filaments protruding from the zeolite micropores. This confirms
that insoluble coke results from the growth of the soluble coke molecules trap-
ped in the pores. Its structure is more like pentacene (linear polyaromatic).

3.3 Location of coke

Whatever the zeolite, most of the soluble coke molecules are too volatile and
too weakly basic to be located on the outer surface. This is quite obvious with
HZSM5, HER1 and HMOR since the major components of the soluble coke
15

have boiling points below the reaction temperature. In the case of I-SHY, the
boiling points of some components are slightly higher than the reaction tem-
perature (Table 3); however their vapor pressure would be high enough for
them to appear in the gas phase if they were located on the outer surface. The
soluble coke molecules are therefore necessarily located in the pores. The size
of the molecules is quite compatible with the size of the cavities (USHY and
HERI), of the channel intersections (HZSM5) or of the channels (HMOR),
However their size is greater than that of the apertures of the cavities (USHY,
HERI) or of the channels (HZSM5). With HMOR we have a special case: the
molecules of the soluble coke which are of a smaller size than the channels and
are highly volatile should diffuse easily out of the zeolite. Therefore they are
probably blocked inside the zeolite by the insoluble coke molecules. The com-
parison between soluble coke molecules and the pore structure of the various
zeolites allows us to imagine how these molecules are located (and trapped) in
the cavities (USHY, HERI), at the channel intersections (HZSM5) or in the
channels between insoluble coke molecules (HMOR).
Except with HMOR, insoluble coke molecules result from the growth of sol-
uble molecules. At least part of each insoluble coke molecule is therefore lo-
cated in the micropores of the zeolite. Here again, given the size of these mol-
ecules, part of each is also located outside the pores as can be seen by electron
microscopy [ 221.

3.4 Mode of coke formation

It is possible in the light of the coke composition to discuss the mode of coke
formation. With all the zeolites coke formation occurs probably through the
same steps (Fig. 10): cracking of n-heptane, oligomerization of the olefinic
cracking products, cyclization of the oligomers, transformation through hy-
drogen transfer into monoaromatics, alkylation of these monoaromatics then
cytilization and hydrogen transfer to give biaromatics, triaromatics etc. Bi-
molecular reactions of naphthenes are improbable because their concentration
is always very low.
All these reactions can be catalyzed by acid sites. Fig. lla gives a possible

nC7 - C3-C4 olefins - oligomers

alkylation I

Polyaromatics c--- Monoaromatics - Naphtenes

cyclization

Hydrogen transfer

Fig. 10. Mode of coke formation during n-heptane (nC,) cracking.


16

al
L

C C-E-C c c c t;:\
c&c - c_;_c_;_c
- &c’_-‘i-c - f
c-c-C-C-E-C
1
,.i.,.,l:,
yJ _ yJ; c- +R+ I - I;;
c-c-c-c-c_c=c

b)

Fig. 11. Coking mechanisms. Modes of formation of (a) toluene from propene and isobutene and
(b) naphthalene from toluene. Hydride shifts are not shown.

mode of formation of a monoaromatic (toluene) from propene and isobutene


(via the cyclization of monoolefinic carbenium ions). Fig. llb shows how a
supplementary aromatic ring could be formed. However reactions through rad-
ical intermediates could also participate to the formation of coke. Thus Sexton
et al. proposed that the external coke formed during the methanol to gasoline
process resulted from the thermal cracking of methanol [ 211.
The pore structure can play a significant role in these various steps:
( 1) Steric constraints can limit the reactions occurring through intermedi-
ates having a size comparable to that of the space available near the acid sites.
Thus the size and the shape of the polyaromatic coke components were deter-
mined by the size and the shape of the cavities or of the channel intersections
(Table 3 ). Owing probably to more pronounced steric constraints in the trans-
formation of butenes through bimolecular reactions (e.g. alkylation, hydrogen
transfer) the olefin-to-alkane ratio in the C, was 1.5 times higher than in the
C, on HZSM5 (narrow channel intersections) while it was 3 times smaller on
USHY (large cavities).
(2) Because of limitations in the diffusion of reactive compounds or of their
trapping (e.g. isobutene in erionite cages) these compounds participate signif-
icantly in the formation of coke.

4. MODE OF DEACTIVATION

The soluble coke components are too weakly basic to be retained adsorbed
on the active acid sites at the high temperature chosen for the reaction. The
deactivation therefore cannot result from a poisoning of the active acid sites
but from the limitation or the blockage of the access of the reactant to these
sites. The deactivating effect of coke depends obviously on its location, e.g. in
the cavities or at channel intersections, or on the outer surface etc., and on the
pore structure of the zeolite. To discuss the mode of deactivation of the zeolites
the following points will be considered:
(1) The deactivating effect of the coke molecules: for this the number of
coke molecules must be determined for the soluble and for the insoluble frac-
tion of coke. For the latter, the average number of aromatic rings of the com-
ponents and consequently their average molar weight M was deduced from the
hydrogen-to-carbon atomic ratio. The estimation of the number of insoluble
coke molecules (from the coke content and from M) is obviously imprecise.
(2) The limitations or the blockages of the access to the pores of various
adsorbates, caused by the coke molecules. The first adsorbate which was cho-
sen was n-hexane, a molecule with a kinetic diameter similar to that of the
reactant n-heptane. However, because of the great difference between the ad-
sorption (273 K) and reaction (723 K) temperatures the conclusions drawn
from n-hexane adsorption cannot apply directly to the reactant n-heptane. It
is why adsorption of a less bulky molecule (nitrogen) will also be considered.

4.1 Deactivating effect of the coke molecules

The deactivating effect of the coke molecules depends very much on the
zeolite (Fig. 12 ): initially one coke molecule has a deactivating effect on USHY
4 times more pronounced than on HZSM5. On HMOR and on HER1 the deac-
tivating effect is still more pronounced: 10 to 20 times more than on HZSM5.
The mode of deactivation differs therefore with the various zeolites.
The number of coke molecules which causes the complete deactivation of
the zeolites was estimated by extrapolation of the curves of Fig. 12 to a zero
activity. This number is compared to the number of strong acid sites A2 (Table
4). While on USHY and on HZSM5 the values are similar, on HMOR and on
HER1 the number of coke molecules is about 7 times smaller than the number
18

1 2 *"k

Fig. 12. Deactivating effect of coke molecules. Change in the residual activity AR versus dnk (10”’
molecules g-i). dn,=difference between the number of coke molecules and this number after 2
minutes reaction. USHY ( l ), HMOR (* ), HZSMS (* ) and HER1 ( A ) .

TABLE 4

Deactivating effect of coke molecules.


(nk) r number of coke molecules for complete deactivation ( 1020g-l), nAJnumber of strong acid
sites (10’” gg’), nAl: total number of acid sites (102” g-l).

(nh)r nA, n.4,

USHY 2.4 3.4 18.5

HMOR 1 7.5 10.8

HZSMS 1.7 1.6 2.2

HER1 1.2 8 12.4

of strong acid sites. AZ. We can therefore conclude that for these last two zeo-
lites deactivation is mainly due to pore blockage. This is confirmed when the
relative decrease in activity l-AK (the residual activity) is compared to the
ratio nk/nAn (number of coke molecules/number of strong acid sites). In Fig.
13, giving l-AR as a function of nJn,,, the experimental values are plotted
and straight lines are drawn with different slopes. If the experimental points
were located on the straight line with the slope equal to 1, it could be concluded
that one coke molecule renders inactive one strong acid site, (with the slope
equal to 2,2 acid sites and so on), provided however that these acid sites have
the same cracking activity. Initially one coke molecule makes inactive over 20
strong acid sites of HER1 or of HMOR which is quite characteristic of pore
blockage. On the contrary on HZSM5 at low coke content 4 coke molecules are
19

l-AR

0.2 0.4 U.b

Fig. 13. Relative decrease in activity (I-A,) as a function of the ratio of the number of coke
molecules to the number of strong acid sites (n,/nA,). Experimental values: USHY (0 ), HZSM5
(I). Straight lines with slope=l, 2,5; HMOR (*), HER1 (A), slope=lO, 20.

needed to deactivat,e only one strong acid site. This low deactivating effect of
coke is characteristic of a limitation of the access to the acid sites. USHY has
an intermediate behaviour. The deactivating effect of coke changes with the
coke content: it decreases on HERI, HMOR and USHY but increases on
HZSM5.

4.2. Limitation or blockage of the access to the pores

Adsorption measurements on the coked zeolites give the value of VA, the
pore volume made inaccessible to the adsorbate. The comparison between VA
and V, the volume really occupied by coke allows to determine the eventual
blockage by coke of the access to part of the pore volume. V, is given by the
ratio of mass of coke deposited per gram of zeolite to density of coke estimated
from its composition. If V,/V, equals 1 there is no pore blockage; if V,/ VA < 1
there is pore blockage.
With HZSM5, VA does not depend on the adsorbate nitrogen (Fig. 14) or
n-hexane (Fig. 15). At low coke content, V,/V, is close to 1 (no pore block-
age). On the contrary a significant pore blockage is found for high coke con-
tent. Thus for 7 wt.-% coke, V,/ VA = 0.3 which means that coke blocks the
access of the adsorbates (and probably of the reactant n-heptane) to a volume
3.3 times greater than the volume it occupies. This pore blockage can be at-
tributed to the insoluble coke molecules which by surrounding the zeolite crys-
tallites [ 22 ] block the access of the reactant to the active sites.
With USHY at low content V, equals half the volume inaccessible to
n-hexane (Fig. 15) and the whole volume inaccessible to nitrogen (Fig. 14).
This can be explained if the size of the coke molecules is considered. Indeed
the first molecules trapped in the cavities (Table 3 for a 2 wt.-% coke content)
have a volume of about 200 A” which corresponds to about a of the volume of
C

0 5 10 15 xc

Fig. 14. Change of the ratio ( V,/V, ) of the pore volume really occupied by coke to the volume
made inaccessible to nitrogen as a function of the coke percentage (% C ) . USHY ( 0 ) , HMOR
(*), HZSMS (*),HERl!-(A).

0 5 10 l!i C

Fig. 15. Change of the ratio ( V,/V,) of the pore volume really occupied by coke to the volume
made inaccessible to n-hexane as a function of the coke percentage ( % C ) . USHY ( l ) , HMOR
(*),HZSMB ($),HERI (A).

the supercage. Consequently, certain coke molecules have no effect on the dif-
fusion of n-hexane (at 473 K) through the supercage while some limit or even
block this diffusion. Moreover the diffusion of nitrogen (kinetic diameter of
3.6 A against 4.3 A for n-hexane or n-heptane) is not inhibited by coke. It is
therefore probable, given the greater mobility of the molecules at the reaction
temperature (723 K), that coke does not block the diffusion of n-heptane
through the supercage. V,/V, decreases when the coke content increases, the
decrease being more pronounced with the adsorbate nitrogen. For coke con-
tents greater than 7 wt.-% VJV, no longer depends on the adsorbate. It is
pore blockage which causes the decrease of V,/V,. At high coke content the
size of the coke molecules (soluble or insoluble) is such that they block the
diffusion (of n-hexane, of nitrogen and also of the reactant n-heptane) through
the supercages in which they are located. It must be noted that V,/ VA remains
21

practically constant for coke contents above 7 wt.-% that is to say when the
increase in the coke is due to the formation of insoluble coke molecules only
(Fig. 9 ) . This means that these coke molecules, although bulkier than the sol-
uble molecules do not create a greater amount of pore blockage. Their shape
(quasi-linear filaments) and their location (protruding from the micropores)
make it possible to understand this observation.
With HMOR and HER1 the volume made inaccessible to the adsorbate
n-hexane is about 10 times greater than the volume really occupied by coke
(Fig. 15). This confirms that the deactivation of these zeolites occur through
pore blockage. With HERI, V,/ VA does not depend on the adsorbate. On the
contrary with HMOR, V,/V, for nitrogen (Fig. 14) is different than for
n-hexane (Fig. 15): in particular this ratio is equal to 1 at low coke contents
which means that all the volume not occupied by coke is accessible to nitrogen.
This is due to the ability nitrogen has to diffuse through the narrow channels
therefore reaching the volume of the large channels located between two plugs
of coke. However V,/V, for nitrogen decreases rapidly and at 4.5 wt.-% coke
is equal to the value found for n-hexane. This blockage of the nitrogen access
to the pore volume is probably due to insoluble coke molecules deposited on
the outer surface.

4.3 Modes of deactivation

The effect the coke molecules have on the cracking activity of the various
zeolites and on their capacity for the adsorption of nitrogen and of n-hexane
can be explained by considering the three following modes of deactivation.
(1) Limitation of the access of the reactant molecules to the active sites.
(2) Blockage of the access to the sites of the cavities (or of the channel
intersections) in which the coke molecules are located.
(3) blockage of the access to the sites of the internal pore volume: (i) chan-
nel blockage or (ii) blockage of the access to the cavities (or channel intersec-
tions) in which there are no coke molecules.
The first mode was found at low coke content on HZSM5. Indeed, the access
to the volume not occupied by coke is not blocked ( VR/VA = l), 4 coke mole-
cules are necessary to deactivate one acid site. The coke molecules trapped at
channel intersections do not block the access of the reactant n-heptane to the
active sites. Actually, the first molecules trapped (C,,H,, alkylbenzenics) can
move in the channel intersection and therefore do not obstruct the access to
the active sites, and also can let the reactant and product molecules diffuse
through the channel intersection. This mode of deactivation occurs probably
with all the zeolites but often at too low coke contents to be observed. In par-
ticular this mode could be the main one responsible for the deactivation of
USHY for coke contents below 2 wt.-%, Indeed the volume inaccessible to
nitrogen (and probably to the reactant n-heptane) is close to V,, the volume
22

actually occupied by coke. However one coke molecule seems to deactivate


about 5 active sites. This high value of deactivating effect, abnormal with this
mode of deactivation, can be attributed to the heterogeneity in strength of the
acid sites. Indeed coke is preferentially formed on the strongest (hence the
most active) acid sites and deactivate preferentially these sites.
With the second mode of deactivation the volume inaccessible to the reac-
tant n-heptane is equal to the volume of the cavities or of the channel inter-
sections in which the coke molecules are located. This is practically what is
found on USHY for coke contents above 2 wt.-%. However, although there is
an average of one strong acid site per supercage, each coke molecule seems to
deactivate more than 2 active sites, Again the heterogeneity in strength of the
acid sites of USHY is responsible for this too-high deactivating effect of coke.
The third mode of deactivation is shown clearly with HMOR and with HERI.
With HMOR the deactivation occurs through blockage of the large channels
by insoluble coke molecules. The reactant has no access to the active sites of
the large channels situated between two insoluble coke molecules. Therefore
the deactivation is very rapid and the volume inaccessible to the adsorbate
n-hexane is much greater than the volume occupied by the coke molecules.
With HER1 the deactivation is due to a blockage of the access of the reactant
to the inner cages first by coke molecules (or precursors like isobutene) formed
in the cages near the outer surface then by coke molecules deposited on the
outer surface.

4.4 Conclusion

The pore structure of zeolites determines for a large part the deactivating
effect of coke.
( 1) When the pore system is constituted of non interconnecting channels
as in HMOR (Fig. 16), deactivation occurs through pore blockage (Mode 3).
The first coke molecules formed in the large channels are retained because of
their low volatility. One coke molecule is enough to inhibit the diffusion of the
reactant to the active sites of the channels. Coke has therefore a great deacti-
vating effect.
(2) When the pore system is constituted of interconnecting channels with-
out cavities like on HZSM5 (Fig. 17) deactivation occurs initially through

Fig. 16. Mode of deactivation of a zeolite with non-interconnecting channels (e.g. HMOR).
23

Fig. 17. Modes of deactivation of a zeolite with interconnecting channels and without cavities (e.g.
HZSM5).

Fig. 18. Modes of deactivation of a zeolite with interconnecting cages with large apertures (e.g.
USHY ).

Fig. 19. Modes of deactivation of a zeolite with cages with small apertures (e.g. HERI).

limitation of the access to the active sites (Mode 1 ), then blockage of the access
to the sites of the channel intersection in which the coke molecules are situated
(Mode 2). Lastly at high coke content coke molecules located on the outer
surface of the crystallites can block the access to the sites of channel intersec-
tions in which there are no coke molecules (Mode 3). It is exactly the same
situation when the pore system is constituted by interconnecting cages whose
apertures are larger than the size of the reactant and product molecules (e.g.
USHY, Fig. 18). For these zeolites coke has a moderate deactivating effect.
(3) When the pore system comprises cavities with small apertures (e.g.
HERI, Fig. 19) molecules of coke precursors limit the access to the active sites
of the cavities where these molecules are trapped and also to the active sites of
the inner cavities. They react rapidly with other compounds to give coke mol-
24

ecules which block the access to the active sites of the inner cages. This deac-
tivation, as in the non-interconnecting channel system, is therefore very rapid.

5. CONCLUSION

Thanks to a combination of different techniques, such as catalytic activity


measurements, extraction of coke from the zeolite and analysis by, e.g. GC,
HPLC, HNMR, MS and so on, and adsorption on the coked zeolites, it has
been possible to understand better the modes of coking and of deactivation of
zeolites. Among these techniques it is the analysis of coke after dissolution of
the zeolite which has given the fundamental information on the nature and
the distribution of the coke components. It must be noted that the techniques
generally used to characterize coke such as in situ spectroscopic techniques
give the chemical identity of the coke components only.
The main conclusions concerning the effect of the pore structure on the
coking and deactivation of the zeolites drawn from this study are:
(1) Coking is a shape-selective reaction: both its rate and its selectivity (na-
ture and distribution of the coke components) depend on the pore structure.
Thus the coking rate is greater when (i ) the space available for the formation
of precursors is larger than their volume and (ii) the coke precursors diffuse
more slowly from the pores to the gas phase. At high reaction temperature the
retention of coke molecules is due to trapping. Therefore their size is inter-
mediate between the size of the apertures and that of the cavities (or of the
channel intersections); their shape is related to the shape of the cavities.
(2 ) The deactivating effect of coke is very pronounced on zeolites having
monodimensional pore structure or large cavities with small apertures.
(3) Deactivation of zeolites by coke does not occur through site poisoning
but through limitation or blockage of the access of the reactant to the active
sites.
However, coking and deactivation do not depend on the zeolite pore struc-
ture only. They depend also on the operating conditions: reaction time, tem-
perature [ 11,29,30,32-34,36,37,43,45,48,49,52,65], pressure and nature of the
reactant [l&31,32,58,66,67]. Some of the above conclusions are clearly not
applicable under different operating conditions. For instance, this would be
the case for the cause of the retention of the coke molecules in the zeolite pore
which is their low volatility at low reaction temperature and their trapping at
high temperature. Moreover coking and deactivation rates as well as coke com-
position depend obviously on the characteristics of the active sites, their
strength and density and so on. It is even sometimes difficult to discriminate
between the effects of the pore structure and of the acidity. That is why inves-
tigations are at present being carried out in our laboratory in order to specify
how exactly the strength and the density of the zeolite acid sites affect the
coking and deactivation phenomena.
25

REFERENCES

J. Barbier, Appl. Catal., 23 (1986) 225.


D. Decroocq, Catalytic Cracking of Heavy Petroleum Fractions, Technip, Paris, 1984.
J.W. Beekman and G.F. Froment, Ind. Eng. Chem. Fundam., 18 (1979) 245.
G.F. Froment, in B. Delmon and G.F. Froment (Eds.), Studies in Surface Science and Ca-
talysis, Vol. 6, Catalyst Deactivation, Elsevier, Amsterdam, 1980, p.1.
L.D. Rollmann, J. Catal., 47 (1977) 113.
L.D. Rollmann and D.E. Walsh, J. Catal., 56 (1987) 139.
P. Magnoux, P. Cartraud, S. Mignard and M. Guisnet, J. Catal., 102 (1987) 242.
E.G. Derouane, in B. Imelik, C. Naccache, G. Coudurier, Y. Ben Taarit and J.C. Vedrine
(Eds.), Studies in Surface Science and Catalysis, Vol. 20, Catalysis by Acids and Bases,
Elsevier, Amsterdam, 1985, p. 221.
9 S. Slinking, A.V. Kucherov, D.A. Kondratyev, T.N. Bondarenko, A.M. Rubinstein and Kh.M.
Minachev, in Y. Murakami, A. Iijima and J.W. Ward (Eds.), New Developments in Zeolite
Science and Technology, Proc. 7th Int. Zeolite Conference, Kodansha, Tokyo and Elsevier,
Amsterdam, 1986, p. 819.
10 P. Dejaifve, A. Auroux, P.C. Gravelle, J.C. Vedrine, Z. Gabelica and E.G. Derouane, J. Catal.,
70 (1981) 123.
11 P. Magnoux, P. Roger, C. Canaff, V. Fouche, N.S. Gnep and M. Guisnet, in B. Delmon and
G.F. Froment (Eds.), Studies in Surface Science and Catalysis, Vol. 34, Catalyst Deactiva-
tion 1987, Elsevier, Amsterdam, 1987, p. 317.
12 B.E. Langner, Appl. Catal., 2 (1982) 289.
13 D.M. Bibby, N.B. Milestone, J.E. Patterson and L.P. Aldridge, J. Catal., 97 (1986) 493.
14 S. Mignard, P. Cartraud, P. Magnoux and M. Guisnet, J. Catal., 117 (1989) 503.
15 J. Volter, V. Ktirschner, J. Caro, J. Kiirger, E. Schreier, B. Parlitz and H.G. Jerschkewitz,
Proc. VIth Int. Symp. Het,erogeneous Catalysis, Sofia, 1987, part 2, p. 112.
16 J. Karger, H. Pfeifer, D. Freude, J. Caro, M. Bulow, G. Ohlmann, in Y. Murakami, A. Iijima
and J.W. Ward (Eds ), New Developments in Zeolite Science and Technology, Proc. 7th Int.
Zeolite Conference, Kodansha, Tokyo and Elsevier, Amsterdam, 1986, p. 633.
17 H. Biilow, J. Caro, J. Viilter and J. Karger, in B. Delmon and G.F. Froment (Eds.) Studies
in Surface Science and Catalysis, Vol. 34, Catalyst Deactivation 1987, Elsevier, Amsterdam,
1987, p. 343.
18 J. Karger, H. Pfeifer, J. Care, M. Btilow, H. Schlodder, R. Mostowicz and J. Viilter, Appl.
Catal., 29 (1987) 21.
19 J. VGlter, J. Caro, M. Biilow, B. Fahlke, J. KBrger and M. Hanger, Appl. Catal., 42 (1988) 15.
20 G.D. MC Lellan, R.F. Howe, L.M. Parker and D.M. Bibby, J. Catal., 99 (1986) 486.
21 B.A. Sexton, A.E. Hughes and D.M. Bibby, J. Catal., 109 (1988) 126.
22 P. Gallezot, G. Leclercq, M. Guisnet and P. Magnoux, J. Catal., 114 (1988) 100.
23 M. Guisnet, P. Magnoux and C. Canaf in New Developments in Zeolite Science and Tech-
nology, Discussion the 7th Int. Zeolite Conf., Japan Association of Zeolite, Tokyo, 1986, p.
111.
24 P.B. Venuto and L.A. Hamilton, Ind. Eng. Chem. Prod. Res. Dev., (1967) 190.
25 B.E. Langner, Ind. Eng. Chem. Process Des. Dev., 20 (1981) 326.
26 L.D. Rollmann and D.E. Walsh, in J.L. Figuereido (Ed.), NATO ASI Series E, Vol. 54, Prog-
ress in Catalyst Deactivation, Martinus Nijhoff Publishers, The Hague, 1982, p. 81.
27 T.J. Weeks Jr., C.L. Angell, I.R. Ladd and A.P. Bolton, J. Catal., 33 (1974) 256.
28 M. Guisnet, P. Magnoux and C. Canaff, in R. Setton (Ed.), NATO ASI Series C, Vol. 165,
Chemical reactions in organic and inorganic constrained systems, Reidel Publishing Com-
pany, Dordrecht, 1985, p. 131.
26

29 H.G. Karge, E.P. Blodingh, J.P. Lange and A. Gutsze, Proc. Int. Symp. on Zeolite Catalysis,
Siofok, 1985, p. 639.
30 J.P. Lange, A. Gutsze, J. Allgeier and H.G. Karge, Appl. Catal., 45 (1988) 345.
31 H.G. Karge and E.P. Boldingh, Catal. Today, 3 (1988) 53.
32 H.G. Karge and E.P. Boldingh, Catal. Today, 3 (1988) 379.
33 D.G. Blackmond, J.G. Goodwin and J.E. Lester, J. Catal., 78 (1982) 34.
34 D.G. Blackmond, J.G. Goodwin and J.E. Lester, J. Catal., 78 (1982) 247.
35 L. Kubelkova, J. Novakova, M. Tupa and Z. Tvaruzkova, in Proc. Int. Symp. on Zeolite
Catalysis, Siofok, 1985, p. 649.
36 A.K. Ghosh and R.A. Kydd, J. Catal., 100 (1986) 185.
37 D. Eisenbach and G. Gallei, J. Catal., 56 (1979) 377.
38 .J. Haber, J. Komorek and T. Romotowski, in Proc. Int. Symp. on Zeolite Catalysis, Siofok,
1985 p. 671.
39 P.E. Eberly, J. Phys. Chem., 71 (1967) 1717.
40 M. Laniecki and H.G. Karge, in Proc. VIth Int. Symp. Heterogeneous Catalysis, Sofia, 1987
Part 2, p. 129.
41 E.G. Derouane, J.P. Gilson and J.B. Nagy, Zeolites, 2 (1982) 42.
42 L. Carlton, R.G. Copperthwaite, G.H. Hutchings and E.C. Reynhardt, J. Chem. Sot. Chem.
Commun., (1986) 1008.
43 S. Maixner, C.Y. Chen, P.J. Grobet, P.A. Jacobs and J. Weitkamp, in Y. Murakami, A. Iijima
and J.W. Ward (Eds,) New Developments in Zeolite Science and Technology, Proc. 7th Int.
Zeolite Conference, Kodansha, Tokyo and Elsevier, Amsterdam, 1986, p. 693.
44 J. Weitkamp and S. Maixner, Zeolites, 7 (1987) 6.
45 H. Lechert, W.D. Basler and M. Jia, Catal. Today, 3 (1988) 23.
46 M. Neuber, S. Ernst, H. Geerts, P.J. Gobet, P.A. Jacobs, G.T. Kokotailo and J. Weitkamp,
in B. Delmon and G.F. Froment (Eds.), Studies in Surface Science and Catalysis, Vol. 34,
Catalyst Deactivation 1987, Elsevier, Amsterdam, 1987, p. 567.
47 A. Gutsze, J.P. Lange, H.G. Karge and J. Allgeier, J. Catal., 113 (1988) 525.
48 J.P. Lange, A. Gutsze and H.G. Karge, J. Catal., 114 (1988) 136.
49 H.G. Karge, J.P. Lange, A. Gutsze and M. Kamecki, J. Catal., 114 (1988) 144.
50 B.E. Langner and S. Meyer, in B. Delmon and G.F. Froment (Eds.), Studies in Surface
Science and Catalysis, Vol. 6. Catalyst Deactivation, Elsevier, Amsterdam, 1980, p. 91.
51 H.S. Bierenbaum, R.D. Partridge and A.H. Weiss, in W.M. Meier and J.B. Uytterhoeven
(Eds.), Advances in Chemistry Series, Molecular Sieves, ACS, Washington, 1973, p. 605.
52 B.E. Langner, Ind. Eng. Chem. Process Des. Dev., 20 (1981) 326.
53 K.W. McLaughlin and R.C. Anthony, AIChE Journal, 31 (1985) 927.
54 H. Schulz, Z. Siwei and W. Baumgartner, in B. Delmon and G.F. Froment (Eds.), Studies in
Surface Science and Catalysis, Vol. 34, Catalyst Deactivation 1987, Elsevier, Amsterdam,
1987 p. 479.
55 M. Guisnet, P. Magnoux and C. Canaff, in Y. Murakami, A. Iijima, and J.W. Ward (Eds.),
New Developments in Zeolite Science and Technolgy, Proc. 7th Int. Zeolite Conference, Ko-
dansha, Tokyo, and Elsevier, Amsterdam, 1986, p. 701.
56 P. Magnoux, P. Cartraud, S. Mignard and M. Guisnet, J. Catal., 106 (1987) 235.
57 P. Magnoux and M. Guisnet, Zeolites, 9 (1989) 329.
58 P. Magnoux, Thesis, Poitiers, 1987.
59 C. Mirodatos and D. Barthomeuf, J. Catal., 93 (1985) 246.
60 C. Mirodatos and D. Barthomeuf, J. Catal., 114 (1988) 121.
61 G. Bourdillon, Thesis, Poitiers, 1985.
62 D.H. Olson, W.O. Haag and R.M. Lago, J. Catal., 61 (1980) 390.
27

63 R.M. Lago, W.O. Haag, R.J. Mikowsky, D.L. Olson, SD. Hellring, K.D. Schmitt and G.T.
Ken, in Y. Murakami, A. Iijima and J.W. Ward (Eds.), New Developments in Zeolite Science
and Technology, Proc. 7th Int. Zeolite Conference, Kodansha, Tokyo, and Elsevier, Amster-
dam, 1986, p. 677.
64 M. Guisnet, F.X. Cormerais, Y.S. Chen, G. P6rot and E. Freund, Zeolites, 4 (1984) 108.
65 P. Magnoux, V. Fouche and M. Guisnet, Bull. Sot. Chim., (1987) 696.
66 D.E. Walsh and L.D. Rollmann, J. Catal., 49 (19’77) 369.
67 D.E. Walsh and L.D. Rollmann, J. Catal., 56 (1979) 195.

You might also like