Vous êtes sur la page 1sur 21

CATALYSIS

A: GENERAL
ELSEVIER Applied Catalysis A: General 133 (1995) 219-239

Review
Alternative processes for the production of styrene
F. Cavani *, F. Trifir6
Dipartimento di Chimica Industriale e dei Materiali, Universit~ di Bologna, Viale Risorgimento 4,
40136 Bologna, Italy

Received 26 June 1995; accepted 1 September 1995

Abstract

T h i s short review c o m p a r e s different technologies for the s y n t h e s i s o f styrene w h i c h are currently


studied as alternatives to the industrial d e h y d r o g e n a t i o n o f ethylbenzene: d e h y d r o g e n a t i o n of ethyl-
b e n z e n e followed by oxidation o f h y d r o g e n , catalytic a n d stoichiometric oxidative d e h y d r o g e n a t i o n
o f ethylbenzene, and d e h y d r o g e n a t i o n in m e m b r a n e reactors. T h e a d v a n t a g e s and d r a w b a c k s o f each
t e c h n o l o g y are illustrated and discussed, and the catalytic s y s t e m s e m p l o y e d are described.

Keywords: Ethylbenzene; Styrene; Dehydrogenation; Oxidative dehydrogenation; Membrane reactor

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
2. Dehydrogenation of etylbenzene . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
2.1. The chemistry of the reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
2.2. Catalyst compositions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
2.3. The role of steam . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
2.4. The role of promoters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
3. Dehydrogenation of ethylbenzene and oxidation of hydrogen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
3.1. Catalyst and reactor configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
3.1.1. Injection of oxygen with the feed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
3.1.2. Injection of oxygen in the effluent of the first dehydrogenation reactor . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
3. 1.3. Injection of oxygen in the effluent of the dehydrogenation reactor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
3.2. The SMART process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
4. Oxidative dehydrogenation of ethylbenzene . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
4.1. Catalytic oxidation with acid oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
4.1.1. The nature of 'active coke' . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
4.2. Catalytic oxidation with redox oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
4.2.1. Catalytic oxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
4.2.2. Non-catalytic oxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233

* Corresponding author. Tel. ( + 39-51 ) 6443680, fax. ( + 39-51 ) 6443682, e-mail cicatal@boifcc.cineca.it

0926-860X/95/$29.00 1995 Elsevier Science B.V. All rights reserved


SSDIO926-860X(95)002 18-9
220 F. Cavani, F. Trifirb / Applied Catalysis A: General 133 (1995) 219-239

5. M e m b r a n e t e c h n o l o g y ...................................................................................... 234
6. C o n c l u d i n g r e m a r k s ........................................................................................ 237
Acknowledgements ........................................................................................... 238
References ..................................................................................................... 238

1. Introduction

Styrene nowadays is produced by two processes: (i) dehydrogenation of ethyl-


benzene and (ii) as a by-product in the epoxidation of propene with ethylbenzene
hydroperoxide and Mo complex-based catalysts. The former process accounts for
more than 90% of the worldwide capacity, which is approximately 13.10 6 t/year.
The latter process is commercialized by ARCO Chemical (formerly Oxirane) and
by Shell. Approximately 1.2. l 0 6 t/year are currently produced with this technol-
ogy.
The problems encountered in ethylbenzene dehydrogenation, as well as the
possible solutions, are similar to those met in the dehydrogenation of alkanes.
Therefore a careful analysis of the alternative solutions, that have been proposed
for the synthesis of styrene, can be useful to evidence the general problems encoun-
tered in the development of alternative technologies for the simple dehydrogenation
and aimed toward the introduction of a double bond in a hydrocarbon molecule.
The main problems in the actual ethylbenzene dehydrogenation process are:
the need for a reactant recycle, owing to the low conversion achieved per pass
( due to thermodynamic limitations);
the need for high steam-to-hydrocarbon ratios;
the high endothermicity of the reaction (AH298 28.1 kcal/mol) ; =

a slight irreversible deactivation of the catalyst (the lifetime is usually about two
years).
The following alternative techniques have been proposed to give a solution to
the above described problems:
oxidative dehydrogenation, in order to realize an exothermic reaction and shift
completely the equilibrium towards the product formation and to carry out the
reaction at lower temperature.
dehydrogenation, followed by oxidation of the hydrogen, in order to furnish the
heat of reaction from the inside, and in some cases (according to the chosen
solution) shift the reaction equilibrium.
membrane catalysis, in order to shift the equilibrium and to carry out the reaction
at lower temperature.
In the first part of this review the main aspects of dehydrogenation technologies
will be described, and in the latter, the alternative technologies will be examined.
F. Cavani, F. Trifirb /Applied Catalysis A: General 133 (1995) 219-239 221

standard free energy, kcal/mol


A C8Hlo --> CeHe+ H2
200 B CsH,o+H2--> CrHe+CH4 20
C CeH,o--> CeHe+C2H4
~ D CsH,o+SH20-'> 8CO+13H2
~ - ~ ~ E C.H,--> 8C+4H~
100 10

'0

-100

-200

0
I

100
I

200
I

300
I

400
I

500
temperature, C
I

600
I

700
t -10

-20

80O

Fig. 1. Standard free energy, referred to as one mole of reactant, for the several reactions involved in the
ethylbenzene dehydrogenation.

2. Dehydrogenation of ethylbenzene
2.1. The chemistry of the reaction
The following reactions are involved in the dehydrogenation of ethylbenzene;
some of these reactions are thermodynamically favored at low temperature, while
others are at higher temperatures.
Direct reactions:
ethylbenzene ~ styrene+H2 (dehydrogenation)
ethylbenzene ~ benzene + C2H 4 (cracking)
ethylbenzene + Hz ~ toluene + CH4 (hydrogenolysis)
ethylbenzene + H20 ~ CO + H2 (steam reforming)
Consecutiue reactions
styrene ~ precursors of coke (oligomerization)
precursors of coke + H20 ~ CO2 + H2 (water gas shift)
precursors of coke ~ coke (dehydrogenation)
Fig. 1 shows the values of the standard free energy of reaction (referred to one
mole of reactant) for the various reactions as a function of temperature.
A reaction network which can explain all the main products is shown in
Scheme 1. Approaching equilibrium the selectivity to styrene decreases, not due to
the occurrence of consecutive reactions but rather due to the fact that the formation
of styrene is inhibited, and the rate of formation of the by-products via parallel
routes becomes more relevant. In fact, at low conversion the selectivity to styrene
approaches 100% with many different catalyst compositions.

2.2. Catalyst compositions


Only one family of catalysts has been proposed, which contain the same main
elements and different promoters whose choice and amount give rise to catalysts
222 F. Cavani, F. Trifirb /Applied Catalysis A: General 133 (1995) 219-239

ethylbenzene = ~ styrene = coke precursors

toluene benzene COx coke


Scheme 1. Reaction network in the dehydrogenation of ethylbenzene.

presenting different activity/selectivity patterns, range of optimal temperature of


operation and of steam-to-hydrocarbon ratio.
The different types of catalysts compositions always contain Fe and K oxides
and, in addition, one or more promoters. A typical range of compositions for the
main elements and for promoters is the following:
Fe203 45-77 wt.-%
K20 10-27%
Cr203 0-3%
Ce203 0-5%
MoO3 0-3%
MgO 0-10%
AlzO3 0-0.1%
VzO5 0-2.5%
CaO 0-2.5%
The surface area of the several compositions is about 2 m2/g, and the operative
conditions are the following: temperature ranging from 540 to 650C, pressure from
subatmospheric to 2 atm and steam-to-hydrocarbon molar ratio from 4 to 20.0.

2.3. The role of steam

Steam is present in excess with respect to ethylbenzene in all commercial proc-


esses; evolution in the last years has been towards the decrease of the steam-to-
hydrocarbon ratio to molar values lower than 6, essentially through modifications
in catalyst compositions. In fact, excess steam is a penalty in energetic costs.
The overall effect of the increase of the steam-to-hydrocarbon ratio is to increase
the activity, the selectivity at the same level of conversion (or the conversion at
the same value of selectivity), the lifetime and stability of the catalyst.
These overall effects are consequences of the following positive and negative
effects:
shift of the equilibrium towards higher conversions of ethylbenzene through a
decrease of the partial pressures of ethylbenzene and of hydrogen;
supply of the heat of reaction; overheated steam is mixed with cool, fresh feed
in order to reach the right inlet temperature;
decrease of the amount of coke or of coke precursors by steam-reforming reac-
tions (the reaction is catalyzed by K2CO3) [ 1,2] ;
consumption of ethylbenzene and of styrene by steam-reforming;
avoiding catalyst over-reduction and deactivation by controlling the valence state
of iron through the equilibrium shown in Scheme 2 (the reoxidation of Fe304
occurs only with oxygen [ 3 ] and not with water) ;
F. Cavani, F. Trifirb /Applied Catalysis A: General 133 (1995) 219-239 223

H20 02
FeO i ~ FesO, ~ ~ Fe203
H2 H2
Scheme 2. Reactions of the variation of the valence state of iron oxide in the reaction medium.

increase of the rate of formation of the active phase, KFeO2, and minimization
of the deactivation rate under reaction conditions [4] ;
interaction of steam with K2CO 3 forming free KOH.

2.4. The role of promoters

Potassium is the main promoter of Fe203; it increases the activity by more than
one order of magnitude, and also slightly increases the selectivity to styrene and
the stability of the catalyst. It has been well established in the last years that the
promotional role of potassium consists of the formation of a surface ternary com-
pound, KFeOz, which constitutes the active phase [2-4]. Another further role
played by potassium is the formation of well dispersed K 2 C O 3, which acts as catalyst
for the gasification with steam of carbonaceous deposits [ 1,2].
KFeO2 is metastable at low temperature and in the presence of oxygen and water
vapor; this is the reason why only recently in-situ characterization techniques
allowed its formation under reaction conditions to be demonstrated unequivocally
[3,4].
The role of the other promoters for the F e / K / O system is to favor the formation
of KFeO2 and stabilize it under reaction conditions at lower steam-to-hydrocarbon
ratios. Cr and A1 are considered to be structural promoters as they can enter in the
F e 3 + compounds; MgO may act as a support not only for KFeO2 but also for K2CO 3
increasing the selectivity and also the stability of the catalyst [2].
Ce oxide increases the activity and Mo the selectivity; therefore the presence of
both promoters is suggested in an improved catalyst composition [ 5 ].

3. Dehydrogenation of ethylbenzene and oxidation of hydrogen

Dehydrogenation/oxidation processes are characterized by the injection of a gas


containing oxygen either in the effluent or in the feed of a dehydrogenation reactor,
in order to catalytically oxidize, either in part or totally, the co-produced or co-fed
hydrogen. Scheme 3 exemplifies the process chemistry for the case of ethylbenzene
dehydrogenation [6].
The claimed advantages of the hydrogen oxidation coupled to the hydrocarbon
dehydrogenation are:
ethylbenzene ~ } styrene + hydrogen
hydrogen + 1/2 o x y g e n b water
ethylbenzene ~ benzene, toluene, methane, ethylene
Scheme 3. Process chemistry for the ethylbenzene dehydrogenation and hydrogen oxidation.
224 F. Cavani, F. Trifirb /Applied Catalysis A: General 133 (1995) 219-239

the internal supply of the heat for the endothermic reaction, thus avoiding costly
reheating supplied by superheated steam or with heat exchangers, and also min-
imizing cracking reactions by shortening inter-reactor lines;
the shift of the dehydrogenation equilibrium by consuming hydrogen, so achiev-
ing higher product yields. Higher yields are obtained both by increasing the
conversion and by increasing the selectivity when moving far from the equilib-
rium.
The oxidation of hydrogen is realized by a suitable oxidation catalyst whose
properties must include:
to be selective only in the oxidation of hydrogen, thus avoiding the oxidation of
the reagent and of the product;
to be stable in the severe reaction conditions, i.e., at high temperature (550-
650C) and in the presence of steam;
to be very active, in order to consume all oxygen at the end of the oxidation zone,
for safety reasons and for the stability of the catalyst in a downstream dehydro-
genation reactor.
This type of process can be applied with dehydrogenating catalysts which are
stable or not poisoned in the presence of steam; in fact this type of process is also
referred to as 'steam oxidation dehydrogenation'. This is the reason why the appli-
cation of this technology is claimed in the dehydrogenation of ethylbenzene using
iron oxide-based catalysts in the presence of steam and in the dehydrogenation of
paraffins only with either spinel-type supports or supports based on alumina doped
with rare earth, which are particularly stable in the presence of steam.
Alternatively, for catalysts which are sensitive to water, the concentration of
steam can be kept low by controlling the amount of added oxygen.

3.1. Catalyst and reactor configurations

Two types of catalyst compositions have been proposed:


dual catalysts, with different catalytic functions, each one performing one reac-
tion (ethylbenzene dehydrogenation and hydrogen oxidation);
single polyfunctional catalyst.
The choice of a unique catalyst for the two reactions may have some advantages
in simplification of catalyst loading and unloading, or in application in moving or
fluidized bed reactors.
For the two solutions different reactor configurations have been proposed:

3.1.1. Injection of oxygen with the feed


In this case a dual bed is employed, either with a mixture of the two types of
catalysts, or with stratified layers; alternatively, the same catalyst is used for both
reactions in one reactor.
This kind of technical solution is exemplified in the patent by Reitmeier et al.
[ 7], who proposed for the dehydrogenation of ethylbenzene to styrene a dual
F. Cavani, F. Trifirb /Applied Catalysis A: General 133 (1995) 219-239 225

catalyst bed, with an oxidation catalyst based on Pt or Pd. The oxidation catalyst is
inserted in the dehydrogenation reactor by mixing or layering the two catalysts, or
by supporting the active metal on the conventional iron oxide-based dehydroge-
nation catalyst.
Similar solutions have also been claimed for the dehydrogenation of alkanes;
Drehman and Walker [ 8 ] proposed a single polyfunctional catalyst with dehydro-
genation and oxidation properties for the dehydrogenation of n-butane in the pres-
ence of steam. The proposed catalyst seems to be a traditional dehydrogenation
one, based on tin and alkali-promoted Pt, supported on a spinel compound. In order
to obtain higher selectivity to olefins and to minimize the oxidation of hydrocarbons,
hydrogen is introduced in the feed.
In patents, assigned to UOP, Imai and Jan [ 9 ] described a polyfunctional catalyst
for paraffin dehydrogenation, constituted of tin/alkali-doped, alumina-supported
platinum. The authors found in pilot plant tests a selectivity of 95% for the H2
oxidation (the balance was made up of oxygen consumed in hydrocarbon oxida-
tion). Besides alkali metals, other promoters were selected including scandium,
yttrium, lanthanum and actinium [ 10]. This type of catalyst is resistant to severe
hydrothermal conditions.
Font Freide et al. [ 11 ] described an oxidation catalyst based on cordierite-
supported platinum, for the dehydrogenation of n-paraffins in the presence of
oxygen. The reaction is carried out at temperatures close to 800C, where dehydro-
genation occurs thermally in the gas phase; the oxidation catalyst supplies the heat
of reaction by oxidizing hydrogen.
Bricker et al. [ 12] claimed a process for the dehydrogenation of ethylbenzene
with a dual catalyst. The oxidation catalyst is still a tin, alkali-doped platinum one,
but special emphasis was given to the property of the support to resist the severe
hydrothermal conditions. The alumina, after the impregnation with tin, is calcined
at a higher temperature than that of dehydrogenation; the resulting support has an
apparent bulk density of 0.2 ml/g, with more than 40% of its pore volume consti-
tuted of pores with a size larger than 150 nm.

3.1.2. Injection of oxygen in the effluent of the first dehydrogenation reactor


In this case an oxidation reactor is inserted between two dehydrogenation reac-
tors, thus supplying the heat for the second dehydrogenation stage.
In patents assigned to UOP, Imai [13,14] proposed a dual catalyst for the
dehydrogenation of ethylbenzene, with the oxidation catalyst in between two dehy-
drogenation stages, or together in the presence of steam. As an oxidation catalyst,
the author claimed a metal of Group VIII, doped with elements of Groups Ia and
IIa with an ionic radius higher than 0.13 nm, in particular either rubidium or cesium
or barium. The properties of the suggested promoters are their low volatility in the
reaction medium and an effect of stabilization induced on alumina, under severe
hydrothermal conditions.
226 F. Cavani, F. Trifirb /Applied Catalysis A: General 133 (1995) 219-239

dehydrogenation reactors

J. . . . I
. . . . . . . . . . . . . . . . . .

_ separation
I I I zone
A !,i Ii A. . . . . ,
I i,i
,- - ; steam r~'~- -,
, super~atorl I' I I I' I ~ , , I Io~oas
steam I i i i i i ~ I |
, t---..4, I I = ?wasteheat I 1--,
steam , ~..J , I I : exche--ers I I styrer~
I I I l
(xYg_e.n)_,L~. _

alr _ . . . . . . _ .

~._,J condensete
compressor ethylbenzene
+ water vapor
Fig. 2. Simplified flow sheet for the UOP process for the dehydrogenation of eythylbenzene and oxidation of
hydrogen (adapted from Ref. [ 151 ).

In other patents from UOP a process with a dual catalyst for the dehydrogenation
of both ethylbenzene and paraffins is claimed, where the oxidation catalyst is not
only inserted between two dehydrogenation reactors, but is also placed before the
first reactor, where hydrogen added to the feed (or coming from the recycle) is
oxidized in order to supply the heat to the first dehydrogenation stage [ 15 ]. Fig. 2
displays a scheme of the reactor configuration. The amount of the oxidation catalyst
is about 30% of the dehydrogenation one and clearly operates with a higher space
velocity. In the case the process is used for paraffin dehydrogenation, the percentage
of water in the stream must not exceed 2% (achieved by controlling the amount of
injected oxygen), in order to avoid the deactivation of the dehydrogenation catalyst.
The claimed oxidation catalyst was Pt, doped with alkali or alkaline earth elements.

3.1.3. Injection of oxygen in the effluent of the dehydrogenation reactor


The effluent of the oxidation reaction is sent directly to the separation zone. In
this case the heat of reaction is supplied to the fresh feed through heat-exchangers
and the problems related to contamination of the dehydrogenation catalyst with
water and oxygen are overcome.
Herber and Thompson [ 16] described for both ethylbenzene and paraffin dehy-
drogenation a dual catalyst-based process. The effluent of the selective oxidation
stage, which is located downstream the dehydrogenation reactor, is sent directly to
the separation unit. Fig. 3 shows a simplified flow-sheet for the ethylbenzene dehy-
drogenation/oxidation process. The claimed oxidation catalyst is a metal of the
Group VIII, with a promoter with ionic radius greater than 0.13 nm.

3.2. The SMART process

The SMART process for the synthesis of styrene combines the Lummus tech-
nology with the UOP concept for oxidative reheating by selective catalytic oxidation
of hydrogen [6]. The flow sheet of the process is shown in Fig. 4; the oxidation
catalyst is inserted in the second and third reactor. Fig. 5 reports the design of
F. Cavani, F. Trifirb / Applied Catalysis A: General 133 (1995) 219-239 227

dehydrogenetion
reactor
fractionating
r- . . . . . . . . . . . .
ate_m__ _ _ -v zone

'-f,~G asas

'i h y d r o c a r b o n s to
, styrene r e c o v e r y
i
i
i
i
i

= oxidation , = water
o x y g e n .~ react or Ic , ' .... "
..... i
-- t i

ethylbenzene

Fig. 3. Simplified flow sheet of the UOP process for the dehydrogenation of ethylbenzene and oxidation of
hydrogen: injection of oxygen in the effluent of the dehydrogenation reactor (adapted from Ref. [ 16 ] ).
h y d r o g e n oxidation

ethylbenzene
dehydrogenation
"~_ . . . . . . . . . . . . . . . . . . . . . . . . . . .

styrene ~ h y d r o g e n oxidation
. . . . . . . . . . . . . . . . . . . . . . . . .
i
hydrogen 4.. . . . . . . . ~ .......... , it P. . . .

t' light ends ~ h y d r o g e n oxidation I, I, i,


i i i I i i i il

:' ' h--rl '"'-':' I


_ separation
i

ethylbenzene

Fig. 4. Simplified flow sheet of the SMART process (adapted from Ref. [6] ).

utlet
steam and air
i
feed i

.,~uent

~oxidation cat.
~dehydrogenet. cat.

dehydrogenation
catalyst
dehydrogenation catalyst
and oxidation catalyst
Fig. 5. Reactor configuration in the SMART process (adapted from Ref. [6] ).

reactor configuration. The gas flows radially from the center, where the oxidation
catalyst is inserted, while in the external part the conventional dehydrogenation
228 F. Cavani, F. Trifirb / Applied Catalysis A: General 133 (1995) 219-239

catalyst is placed. The two beds of catalysts are separated by screens. The removal
of hydrogen formed in the first reactor shifts the equilibrium to the formation of
higher yields of styrene in the second one. With three reactors, a conversion per
pass higher than 80% can be obtained, at the same styrene selectivity as for the
conventional process. It is interesting to observe that these values are similar to the
best ones obtained in the oxidative dehydrogenation of the ethylbenzene with a
carbon molecular sieve [17].
The SMART process can be used to revamp conventional dehydrogenation
plants, in order to increase the capacity with low investment costs. An example has
been given by Romatier et al. [6] ; the revamping of an existing plant producing
272 000 t/year of styrene has been increased in capacity to 400 000 tpa by adding
two new reactors in a three-stage dehydrogenation process; the existing reactors
are not modified, and the new reactors are added in series. The feed rate can be
increased by 10%, in view of the increased ethylbenzene conversion, and corre-
spondingly the reduced recycle stream.

4. Oxidative dehydrogenation of ethylbenzene

In the oxidative dehydrogenation of ethylbenzene the strong exothermic reaction


is used:
C6Hs-C2H5 + 0.502 --") C6Hs-C2H3 + H20 A/~298= - 29.7 k c a l / m o l
in order to reach the following objectives:
to obtain styrene at almost complete conversion in order to decrease the sepa-
ration costs (dehydrogenation processes operate at conversions lower than 65%) ;
to eliminate or strongly decrease the use of superheated steam (steam/hydro-
carbon = 5/1 in dehydrogenation) ;
to reach higher selectivity (90% in dehydrogenation), not only to minimize the
waste of ethylbenzene but also to simplify the removal of the heat of reaction
and avoid total consumption of oxygen. The transformation of one mole of
ethylbenzene to styrene requires 0.5 02 while the combustion to CO2 requires
13 02.
Belomestnykh et al. [ 18 ] report a reaction pattern with most of the representative
by-products (Scheme 4). Side reactions occur by cracking and by oxygen insertion
reactions on ethylbenzene and on styrene.
In order to minimize the oxygen insertion reactions and to eliminate the problems
which arise in the presence of molecular oxygen (mixture flammability, run-away)
three types of solutions have been proposed:
catalytic oxidation in the presence of molecular oxygen with oxides possessing
medium strength acidity and with nil or limited redox properties [ 19];
oxidation with redox-type oxides in the absence of molecular oxygen [ 20];
electrochemical oxidation [ 21 ].
F. Cavani, F. Trifirb /Applied Catalysis A: General 133 (1995) 219-239 229

ethylbenzene = styrene = styreneoxide


/
acetophenone
a-methylbenzylalcohol \
b benzaldehyde

~ benzene~
carbOnoxides
Scheme 4. Reaction pattern in oxidative dehydrogenation of ethylhenzene (from Ref. [ 18] ).

4.1. Catalytic oxidation with acid oxides

The best performances of mixed oxide-based catalysts investigated after 1981


are reported in Table 1. Previous references from patent and scientific literature
are reported in the paper of Emig and Hofmann [22]. The different oxides are
active in the temperature range 450 to 600C, and operate with a ratio O2/EB 0.8-
2, a concentration of EB in the range 8-15% and a contact time in the range 0.2-4
gear/(gEB h ) .
The best results, among all the catalysts listed in Table 1, have been obtained
with carbon molecular sieve AX21 (Anderson Dev. Co.), at much lower temper-
ature (350C), with 80% conversion and 90% selectivity [ 17]. However, these
results were not confirmed in a recent paper by the Drago and Jurczyk [23].
Comparable performances are obtained by metal phosphates, but at much higher
temperatures [24].
On the basis of the first suggestions of Alkhazov et al. [32], up to the recent
short review of Vrieland and Menon [ 19], it seems well documented and proved
that the active catalyst in ethylbenzene oxidative dehydrogenation is the 'active
coke' which forms in the first hours of reaction and reaches a stationary amount on
the surface of the oxides.

Table 1
Catalytic performance of various catalysts in the oxidative dehydrogenation of ethylbenzene to styrene

Catalyst T (C) Conversion (%) Selectivity (%) Ref.

SnO2-P205 450 38 82 [ 25 ]
Zr phosphate 450 55 86 [ 22 ]
SIO2-A1203 450 62 71 [ 26]
Clinoptilolite 475 59 85 [ 27 ]
Na sodalite 475 44 85 127 ]
Pr/Mo/AI/O 500 67 86 11281
Ce4 (P207) 3 550 71 89 [ 24 ]
Al(PO3)3 530 72 91 1241
Ce phosphate 605 76 90 [ 24 ]
Ge phosphate 540 54 64 [ 29 ]
Zr/Sn phosphate 500 64 83 [ 30]
Carbon molsieve CMS AX21 350 80 90 [ 17]
Carbon molsieve Ambesorb 575 300 82 73 [31 ]
230 F. Cavani, F. Trifirb/Applied Catalysis A: General 133 (1995) 219-239

Therefore, the investigated oxides are not catalysts, but rather 'carriers of the
active component'. The formation of the 'active coke' in both internal volume and
on external surface of the different oxides explains the low differences in catalytic
performance observed for the different classes of metal oxides and phosphates, as
shown in Table 1, as well as the higher yield in styrene obtained at much lower
temperature with molecular sieve carbon. With metal oxide carriers higher tem-
peratures are most likely necessary to form and to maintain a certain stationary
amount of 'active coke'.
In the case of 'oxides as carriers' usually conversions not higher than 70% have
been reported. This limitation on conversion is essentially due to the fact that the
conversion of oxygen reaches 100%. In fact, the oxygen-to-styrene ratio in most
publications and patents is not higher than 1, with a preferred concentration of
ethylbenzene of 10%. The choice of this composition, the best for achieving a good
selectivity, requires a costly dilution of air with nitrogen or steam and also causes
oxygen starvation at high conversion. However, it is the preferred composition due
to safety reasons, in order to operate outside the flammability region. The lower
and higher limits of ethylbenzene in air at 30C (those for styrene are similar) are
1.0 and 6.7%, respectively; therefore, operation with 10% ethylbenzene in air would
be unsafe, not at the inlet of the reactor but rather inside it, when the composition
of the reacting mixture enters the flammability zone. The choice of 10% oxygen
with about 10% ethylbenzene gives safe conditions in all sections of the plant
layout.
In order to increase the conversion to values which are economically competitive
with dehydrogenation an oxide carrier presenting higher selectivity must be devel-
oped.
Another factor which can contribute to the non-complete conversion obtained
with the oxide carriers is the strong adsorption of styrene which negatively influ-
ences the rate of transformation of ethylbenzene, as can be deduced from the rate
equations developed by Schraut et al. [ 33 ]. A modification of the catalyst aimed
at facilitating the styrene desorption can contribute to increase the yield to styrene.

4.1.1. The nature of 'active coke'


The composition of coke varies in the first hours of reaction reaching similar C -
H-O compositions (for different carriers). It has been found [ 34] that the 'active
coke' on the surface of Zr-phosphate, under stationary conditions, presents the
following composition: C19H703. Other authors [ 35 ], when using as the carrier a
boron-doped alumina, have found an active coke with composition C22.5H6.603 .
Echigoya et al. [26] found an higher activity and selectivity on mixed oxides
carriers based on silica, where a coke with a H / C ratio less than 1 was formed.
C = O bonds have been identified by various authors [34,36], and radical-like
species were also found [34,35,37]. According to Schraut et al. [34], the radical
species can be attributed to the aroxyl groups shown in Scheme 5. By means of
F. Cavani, F. Trifirb /Applied Catalysis A: General 133 (1995) 219-239 231

O.
Scheme 5. Aroxyl groups in 'active coke' deposited on acid oxides (from Ref. [33] ).

SIMS analysis it was found that the deposited coke is similar to anthraquinone
[34,36].
The activity of coke in the oxydehydrogenation of ethylbenzene has been attrib-
uted to the redox couples which form on the edge of its graphitic structure, after
interaction with oxygen. In the absence of molecular oxygen, 'active coke' [ 22,28 ]
and carbon molecular sieves [17,31] are practically inactive and do not form
hydrogen. A decrease of the oxygen-to-ethylbenzene ratio in feedstock leads to a
decrease of ethylbenzene conversion and to an improvement of styrene selectivity
[ 23 ]. It has been hypothesized that these redox couples consist of quinone/hydro-
quinone and aroxyl/phenol groups. A representation of the mechanism proposed
by Emig and Hofmann [22] is given in Scheme 6; styrene forms after a concerted
mechanism of hydrogen abstraction while carbon oxides partially form from the
combustion of coke.
The different selectivities found on different catalysts, notwithstanding the pos-
session of comparable amounts of coke, has been attributed to the different surface
properties of coke. The most selective coke is that one that is less oxidizable by
oxygen, while the least selective coke is that one that is easily burnt to carbon
oxides even in the presence of high amounts of ethylbenzene [ 38].
The carbon molecular sieve AX21 investigated by Grunewald and Drago [ 17]
presents a surface area of approximately 3000 ~ g , with an oxygen content of

COx active coke ' ~ H204r'~ 1/2 02


Scheme 6. Mechanism ofethylbenzeneoxidative dehydrogenation over 'active coke' (from Ref. [22] ).
232 F. Cavani, F. Trifirb /Applied Catalysis A: General 133 (1995) 219-239

Table 2
Performance of carbonaceous catalysts in the oxidative dehydrogenation of ethylbenzene [ 17,31 ]

Catalyst Surface area (mZ/g) T (C) Conversion (%) Selectivity (%)

PPAN 1" 8 350 11.6 90.5


PPAN 2 50 350 22.4 91.5
PPAN 3 10 350 14.6 76
AC ~ 800 350 26.8 81.7
CMS AX21 c 3000 350 80 90
Ambesorb 563 ~ 540 300 68.3 88
Ambesorb 572 1060 300 74.7 89
Ambesorb 575 785 300 82.3 73
Ambesorb 348F 725 300 55.7 88

Pyrolyzed acrylonitrile, Aldrich.


b Activate carbon, Mallinckrodt.
c Carbon Molecular Sieve AX21, Anderson Dev. Co.
d Carbon Molecular Sieve, Supelco.

6%. It likely that the chemical nature of this oxygen is similar to the quinone-type
proposed to form in active coke when this is supported on oxide carriers. The higher
activity of molecular sieve can be attributed to its high surface area, but the surface
area cannot be the only parameter. In fact the activity data reported in Table 2, for
different types of carbonaceous materials [ 17,31 ], cannot be interpreted only on
the basis of differences in surface area. This is also confirmed by the fact that the
micropores blocking during the reaction, due to polymerization of styrene and
thermal decomposition to coke, did not remarkably affect the catalyst activity [ 23 ].
Apparently, most of the reaction occurs in mesopores and in macropores, as well
as in the external surface area.
The data on PPAN, AC and AX21 were collected after 20 h time-on-stream
[17]. In particular, AX21 was stable after one week on stream. The data on
Ambesorb [ 31 ] were collected after only 1 h of reaction; longer periods of reaction
deactivated the catalyst very quickly.

4.2. Catalytic oxidation with redox oxides

Oxidation of ethylbenzene to styrene has been achieved with redox-type cata-


lysts, based essentially on MgO/V205 and using two different reactor configura-
tions:
catalytic oxidation with air in the presence of steam [ 18,39] ;
oxidation of ethylbenzene with oxygen from the catalyst and reoxidation of the
reduced catalyst with air [20].

4.2.1. Catalytic oxidation


In catalytic oxidation a maximum in activity and selectivity has been observed
at low V205 concentration, as shown in Fig. 6 [ 18]. Belomestnykh et al. [ 18]
obtained the best performance (72% conversion and 90% selectivity) with the 12%
F. Cavani, F. Trifirb /Applied Catalysis A: General 133 (1995) 219-239 233
ethylbenzene conversion, %

I
40~ "
20-

0 I I | 1
0 20 40 60 80 100
vanadia content, wt.%
Fig. 6. Ethylbenzene conversion as a function of vanadia content in the catalyst (from Ref. [ 18] ).

V205 and some unspecified dopant, at the following conditions: temperature 480C,
feed composition E B / O J H 2 0 / N 2 1/ 1/8/20 molar ratio.
Other authors [ 39] also found the highest value of activity and selectivity at the
9% V205, but these values remained unchanged up to 14% of vanadia content. The
best performance was obtained at 520C with a conversion of 65 % and a selectivity
of 93.5%; residence time was 0.4 s and feed composition was EB/O2/H20/N2 1/
1/8/20 (molar ratio).
The performances obtained with redox catalyst are very similar to those obtained
with acid oxides. The only difference is the presence of an higher amount of steam
when redox oxides are used, in order to suppress bulk combustion.
According to some authors [ 18,39] the active sites for the selective oxydehy-
drogenation are surface clusters of V 5+ and V 4 in octahedral coordination. At
low vanadium concentration isolated vanadium species form, in tetrahedral and
octahedral coordination, while when the vanadium content is increased the amount
of vanadium in octahedral coordination increases, with formation of associated
vanadium species, and the catalytic activity is correspondingly increased. For even
higher concentrations of vanadium a Mg-vanadate phase forms and the activity
drops. Hanuza et al. [39] suggested that the sites which abstract hydrogen from
ethylbenzene are surface V 5 = O bonds, while V 4 species dissociate molecular
oxygen.

4.2.2. Non-catalytic oxidation


In non-catalytic oxidation [ 20] mixed oxides with composition ranging from 40
to 60 wt.-% MgO, from 20 to 40 wt.-% SiO2 and from 10 to 30 wt.-% V205 are
used as stoichiometric oxidants of ethylbenzene. The preferred composition is 50
wt.-% MgO, 30 wt.-% SiO2 and 20 wt.-% VzOs. The preferred preparation is spray-
drying of a MgO-SiO2 suspension followed by impregnation of ammonium
metavanadate, preferably with a further amount of silica and calcination at tem-
peratures between 500 and 800C.
234 F. Cavani, F. Trifirb /Applied Catalysis A: General 133 (1995) 219-239

In pulse reactor tests an ethylbenzene conversion of 98.1% and a selectivity to


styrene of 92% were obtained at 505C. After one regeneration with oxygen a small
decrease in activity was observed. Tests in fluidized-bed reactor, at a WHSV of 3.4
h - ' and at the temperature of 500C, gave ethylbenzene conversion of 63% and
selectivity of 83.8%. The regeneration was carried out with a stream containing
10% oxygen in nitrogen. The stripping was carried out with pure nitrogen. These
results, achieved at comparable contact time, are worse than those obtained with
acid oxides as carriers.
The high activity and selectivity obtained in pulse reactor evidences the high
specificity in hydrogen abstraction of transition-elements redox couples. Very
likely, the low performance obtained in flow conditions might be due to an unfa-
vorable formation of carbonaceous deposits. This allows us to exclude coke as the
active component in these catalysts.

5. Membrane technology

Membrane technology has been proposed to modify dehydrogenation and oxi-


dative dehydrogenation processes along the following lines:
to separate hydrogen from the product stream of a dehydrogenation reaction
occurring on one side of the membrane and pure hydrogen is obtained on the
other side of the membrane.
to separate hydrogen from the product stream and to make it react on the other
side of the membrane with an oxygen-containing gas.
to let oxygen diffuse selectively through a membrane in order to make it oxidize
either the hydrogen formed or directly the hydrocarbon, on the other side of the
membrane.
The advantages which are claimed to be achieved when dehydrogenation reac-
tions are carried out inside a reactor which contains a membrane are the following:
to shift the dehydrogenation equilibrium and increase the conversion to the
product with also the possibility to operate at lower temperature so as to decrease
side reactions;
to develop a process in autothermal regime by coupling the endothermic dehy-
drogenation reaction (which occurs on one side of the membrane) with an
exothermic reaction on the other side (such as hydrogenation or hydrogen oxi-
dation) ;
to carry out an oxidative dehydrogenation without mixing the hydrocarbon and
the oxygen, thus minimizing undesired oxygen insertion reactions and avoiding
problems related to flammability of the mixtures and to run-away.
Two types of reactor configuration have been proposed:
F. Cavani, F. Trifirb /Applied Catalysis A: General 133 (1995) 219-239 235

ethylbenzene sweep gas


catalytic
membrane
Catalytic
"~ : : ::~ Membrane
Reactor
permeate
feed side ~ , ~ ; , . ~ side
styrene ~, sweep gas~,
hydrogen

direction across the membrane

ethylbenzene sweep gas


E

1 io..meob .... I ,ne,


~'~'~;~ Membrane
::i~. .~i~ Catalytic
I;~;!iii ~ ; ~ 1 Reactor

styrenel, sweep gas].


hydrogen "
Fig. 7. CMR and ICMR configurations.

the reaction is carried out completely in a membrane reactor;


the membrane reactor is added downstream to a normal reactor, after the reaction
has approached equilibrium conversion (hybrid system); this solution may be
considered as a retrofitting of existing dehydrogenation plants.
Two types of membranes have been proposed:
non-porous (dense) membranes: metals or alloys (Pd; Pd-Ag) [40] ;
porous membranes: Vycor glass, alumina, ceramics or SiC [41,42].
Porous membranes have been investigated more for dehydrogenation reactions
than dense metallic membranes. The effective permeability for porous membranes
is dependent on the pore size, porosity and tortuosity of membrane structure, on
the molecular weight of reagent and products, and on the temperature of separation.
The permeation occurs by Knudsen diffusion and depends on collisions between
the molecule and the walls of the pores with diameters below 10 nm. The separation
factor is inversely proportional to the square root of molecular weight, and therefore
it is very low for C2-C4 paraffins/hydrogen mixtures, while it has a reasonable
value of 7.3 for ethylbenzene/hydrogen.
Two types of catalyst configuration have been proposed: the catalytic membrane
reactor (CMR) and the inert membrane catalytic reactor (IMCR) [43,44]. Fig. 7
shows a scheme of the CMR where the membrane is impregnated with the catalytic
active components, and the IMCR configuration, where the pellets of catalyst fill
the reactor in the feed side, and where the membrane constitutes the wall of the
reactor.
Bitter [ 45 ] has investigated ceramic membrane reactors to separate the hydrogen
from the dehydrogenated products. Fig. 8 shows the reactor configuration: the
catalyst, represented by the shaded area, is in a first reactor zone, as well as in the
membrane reactor and in a post-dehydrogenation reactor zone. The first reactor is
necessary to obtain hydrogen and decrease the amount of hydrocarbon which can
236 F. Cavani, F. Trifirb /Applied Catalysis A: General 133 (1995) 219-239

second dehydrogenation
first dehydrogenation reactor: multitubular third dehydrogenation
reactor membrane reactor reactor

ceramic membrane
permeable to hydrogen retentate outlet,"
\ _ I ethylbenzene, styrene
I '-' IE ;i " I7 - - * H2,
dehydrog. ' .......... [e~'rens
i! ;r !i
ca~

! i ........................
ir?i !~i~ }:2i)i}

~:. S 7 : "~ j f p e r m e a t e outlet: l


A
i hydrogen , t \
...... i v (ethylb.,sreoo,' . . . . . \
ethylbenzene
I dehydrog, dehydrog,
catalyst catalyst
styrene, hydrogen

Fig. 8. CMR: reactors configuration in the process for hydrocarbons dehydrogenation developed by Shell (adapted
from Ref. [44] ).

permeate with hydrogen in the membrane reactor. Hydrogen, with small amounts
of hydrocarbons, permeates through the membrane walls. The post-dehydrogena-
tion reactor treats the hydrogen-rich permeate stream, in order to transform the
reagents which have permeated together with hydrogen through membrane walls.
The latter consist of a layer of aluminum oxide membrane (pore size 10 nm,
thickness 4-10/zm) on a porous y-alumina support (pore size 5-10 nm, thickness
4 mm).
In ethylbenzene dehydrogenation, with a F e / K / V / L i / C r / O catalyst, at a tem-
perature of 625C and with a LHSV of 0.65 h-~, a conversion of 65% and a
selectivity of 94% have been obtained, while without membrane the conversion is
only the 50.7%. A ceramic membrane has been investigated by Wu and Liu [46]
to dehydrogenate ethylbenzene and separate hydrogen in order to increase the yield
in styrene. They proposed an hybrid system where a packed-bed reactor, filled with
a Fe/K oxides-based catalyst, is followed by a membrane reactor also filled with a
dehydrogenation catalyst. The hybrid solution has been proposed in order to avoid
the permeation of ethylbenzene when the concentration of hydrogen is low.
Moser et al. [47] have investigated the dehydrogenation of ethylbenzene with
an alumina membrane reactor, in both CMR and IMCR systems. In the latter
configuration, a conventional dehydrogenation catalyst has been located inside the
tube of porous alumina on the feed site. In the CMR configuration, the pores of the
membrane were first filled with a solution containing the catalyst, in order to achieve
a dispersion of the active components. The conversion of ethylbenzene was
increased above the equilibrium value by 10%, with the IMCR, and by 20-23%,
with the CMR configuration.
Oxygen-permeable membranes have been used in catalytic and electrocatalytic
membrane reactors, for the dehydrogenation with oxygen, hydrogen oxidation and
for the oxidative dehydrogenation of ethylbenzene. Michaels and Vayenas [21 ]
F. Cavani, F. Trifirb /Applied Catalysis A: General 133 (1995) 219-239 237

Table 3
A comparison of the best results reported for the different methods of styrene synthesis from ethylbenzene

Configuration Conversion (%) Selectivity (%) T (C) Catalyst Status

Dehydrogenation 60-65 90 600-650 Fe/K/Cr/O industrial


Dehydrogenation/H2 80 90 SMART technology commercial
oxidation
Oxidative dehydrogenation 76 90 605 Ce phosphate research
Oxidative dehydrogenation 80 90 350 carbon molsieve research
Oxidative dehydrogenation 72 90 480 V/Mg/O research
Non-catalytic oxidation 98 92 505 V/Mg/Si/O research
Membrane technique 65 94 625 Fe/K/V/Li/Cr/O research

studied the vapor phase electrochemical oxidative dehydrogenation of ethylbenzene


to styrene on a polycrystalline Pt electrocatalyst in a stabilized ZrO2 electrochemical
reactor. The dehydrogenation rate was found to be enhanced by a moderate current
density.
Standard Oil [48,49] claims the use of solid multicomponent membranes that
consist of a monophasic mixed metal oxide with a perovskite structure with both
electron- and oxygen-conductive properties. The membrane is impervious to gases
and physically separates two zones in the electrochemical reactor: in the first zone
the oxygen-containing gas is passed and contacted with one surface of the
membrane (the cathode), where molecular oxygen is reduced. The resulting ion is
transported to the other surface of the membrane (the anode), where it meets the
hydrocarbon to be dehydrogenated. Reaction between ionic oxygen and the hydro-
carbon occurs at the surface and releases electrons which are transported to the
cathode surface. A catalyst can be placed in the hydrocarbon zone, to catalyze the
selective oxidation of the hydrogen formed. The process offers the advantage of
overcoming the thermodynamic limitations of direct dehydrogenation, thus achiev-
ing high conversion and limiting the production of carbon oxides which are formed
by direct contact between the oxygen and the hydrocarbon. Single-phase mem-
branes with perovskite structure claimed in the patents are: LaCoOx, Lao.6Sro.4CoO x,
Lao.2Sro.8CoOx, YBa2Cu3Ox. These systems were reported to be effective in the
synthesis of styrene from benzene/ethane and benzene/ethylene mixtures. The
benzene conversion at 700C was 11%, with a selectivity to styrene which
approached 70%, the by-products being biphenyl, ethylbenzene and toluene.

6. Concluding remarks

Table 3 summarizes the best catalytic performances obtained with the different
technologies; dehydrogenation is the only process widely used at a commercial
level. The SMART technology is commercially available, but to our knowledge it
has found .no industrial application till now. The other methods of synthesis are at
a research level. Oxidative dehydrogenation gives better performances in terms of
238 F. Cavani, F. Trifir6 /Applied Catalysis A: General 133 (1995) 219-239

conversion and selectivity than the industrial dehydrogenation process. The best
performance in terms of conversion/selectivity are claimed with the non-catalytic
oxidation on redox oxides, where 92% of selectivity were obtained at 98% conver-
sion. However, for these two alternative technologies the lifetime of catalysts has
not yet been evaluated and therefore pilot plant studies must be carried out in order
to obtain informations about the modifications of catalytic performances with time-
on-stream and about the productivity achievable. It is also necessary to gain a more
quantitative knowledge of all the by-products. In fact, even if present in traces, the
nature and amount of by-products can determine the final success of these new
technologies.

Acknowledgements

This work was sponsored by MURST (Ministero dell' Universit~ e della Ricerca
Scientifica).

References

[ 1] T. Hirano, Appl. Catal., 26 (1986) 65.


[2] D.E. Stobbe, F.R. van Buren, A.J. van Dillen and J.W. Geus, J. Catal., 135 (1992) 533; 548.
[3] M. Muhler, J. Schiitze, M. Wesemann, T. Rayment, T. Dent, R. Schl6gl and G. Ertl, J. Catal., 126 (1990)
339.
[4] T. Hirano, Appl. Catal., 26 (1986) 81.
[5] T. Hirano, Appl. Catal., 28 (1986) 119.
[ 6 ] J. Romatier, M. Bentham, T. Foley and J.A. Valentine, Proc. Dewitt Petrochem. Rev., Houston, Texas, 1992,
p. K1.
[7] R.E. Reitmeier, F.D. Mayfield and J.H. Mayes, US 3 437 703 (1969).
[8] L.E. Drehman and D.W. Walker, US 3 670 044 (1972).
[9] T. Imai and D.Y. Jan, US 4 788 371 (1988).
[ 10] T. Imai and R.J. Schmidt, US 4 886 928 (1989).
[ 11 ] J.J. Font Freide, M.J. Howard and T.A. Lomas, Eur. 332 289 (1989).
[ 12] J.C. Bricker, T. Imai and D.E. Mackowiak, US 4 717 779 (1988).
[ 13] T. Imai, US 4 418 237 (1983).
[ 14] T. Imai, US 4 435 607 (1984).
[ 15] B.V. Vora, US 4 376 225 (1983).
[ 16] R.R. Herber and G.J. Thompson, US 4 806 624 (1989).
[ 17] G.C. Grunewald and R.S. Drago, J. Molec. Catal., 58 (1990) 227.
[ 18] I.P. Belomestnykh, E.A. Skrigan, N.N. Rozhdestvenskaya and G.V. Isaguliants, Stud. Surf. Sci. Catal., 72
(1992) 453.
[ 19] G.E. Vrieland and P.G. Menon, Appl. Catal., 77 ( 1991 ) 1.
[20] L.F.L. Delorme, F. Martins Mendes Cerejo and J.F. Grootjans, Eur. 403 462 (1990).
[21 ] J.N. Michaels and C.G. Vayenas, J. Catal., 85 (1984) 477.
[22] G. Emig and H. Hofmann, J. Catal., 84 (1983) 15.
[23] R.S. Drago and K. Jurczyk, Appl. Catal. A, 112 (1994) 117.
[24] G.E. Vrieland, J. Catal., 111 (1988) 1.
[25] Y. Murakami, K. Iwayama, H. Uchida, T. Hattori and T. Tagawa, Appl. Catal., 2 (1982) 67.
[ 26] E. Echigoya, H. Sano and M. Tanaka, in Proc. 8th Int. Congress on Catalysis, Vol. V, DECHEMA, Frankfurt,
1984, p. 623.
F. Cavani, F. Trifirb /Applied Catalysis A: General 133 (1995) 219-239 239

[27] D.B. Tagiyev and K.M. Minachev, Stud. Surf. Sci. Catal., 28 (1986) 981.
[28] J.J. Kim and S.W. Weller, Appl. Catal., 33 (1987) 15.
[29] M. Turco, G. Bagnasco, P. Ciambelli, A. La Ginestra and G. Russo, Stud. Surf. Sci. Catal., 55 (1990) 327.
[30] G. Bagnasco, P. Ciambelli, M. Turco, A. La Ginestra and P. Patrono, Appl. Catal., 68 ( 1991 ) 69.
[31 ] P. Ciambelli, R.P. Dario and M. Turco, in Proc. 9th Int. Conf. Zeolites, Montreal, Canada, 1992, FP9.
[32] T.G. Alkhazov, A.E. Lisovskii, M.G. Safarov and A.M. Dadasheva, Kinet. Catal., 13 (1972) 509.
[33] A. Schraut, G. Emig and H. Hofmann, J. Catal., 112 (1988) 221.
[ 34] A. Schraut, G. Emig and H.G. Sockel, Appl. Catal., 29 (1987) 311.
[ 351 R. Fiedorow, W. Przystajko, M. Sopa and I.G. Dalla Lana, J. Catal., 68 ( 1981 ) 33.
[36] L.E. Cadus, O.F. Gorriz and J.B. Rivarola, Ind. Eng. Chem. Res., 29 (1990) 1143.
137] L.E. Cadus, L.A. Arrua, O.F. Gorriz and J.B. Rivarola, Ind. Eng. Chem. Res., 27 (1988) 2241.
[ 38] G.E. Vrieland, J. Catal., 111 (1988) 14.
[39] J. Hanuza, B. Jezowska-Trzebiatowska and W. Oganowski, J. Mol. Catal., 29 (1985) 109.
[40] J. Shu, B.P.A. Grandjean, A. van Neste and S. Kaliaguine, Can. J. Chem. Eng., 69 ( 1991 ) 1036.
[41 ] S. Ilias and R. Govind, in R. Govind and N. Itoh (Editors), Membrane Reactor Technology, AIChE Symp.
Series, Vol. 85, No. 268, American Institute of Chemical Engineering, New York, 1989, p. 18.
[42] H.P. Hsieh, AIChE Symposium Series, 268, 1989, p. 53.
[43] Y.M. Sun and S.J. Khang, Ind. Eng. Chem. Res., 27 (1988) 1136.
[44] Y.M. Sun and S.J. Khang, Ind. Eng. Chem. Res., 29 (1990) 232.
[45] J.G.A. Bitter, UK 2 201 159, (1986).
[46] J.C.S. Wu and P.K.T. Liu, Ind. Eng. Chem. Res., 31 (1992) 322.
[47] W.R. Moser, Y. Becker, A.G. Dixon and Y.H. Ma, in Proc. Fifth Annu. Meet. North Am. Membrane Soc.,
Lexington, Kentucky, 1992, 11E.
[48] T.L. Cable, J.G. Frye, W.R. Kliewer and T.J. Mazanec, Eur. 438 902 A2 ( 1991 ).
[49] T.J. Mazanec and T.L. Cable, UK 2 203 446 (1988).

Vous aimerez peut-être aussi