You are on page 1of 45

Reactive Polymers, 20 (1993) 1-45

Elsevier Science Publishers B.V., Amsterdam

State-of-the-Art Report
Cationic ion exchange resins as catalyst
Animesh Chakrabarti and M.M. Sharma *

Department of Chemical Technology, University of Bombay, Matunga, Bombay-400 019, India


(Received November 25, 1992; accepted February 2, 1993)

Abstract

This paper presents a review of the literature pertaining to catalysis by cation exchange
resins. Emphasis has been given to the literature that has appeared during the last seven
years. A variety of reactions catalyzed by cation exchange resins have been tabulated
according to their broad classification; industrially important reactions have been delineated. Examples, where the use of ion exchange resin catalysis has led to an improvement
in the selectivity of the desired products, have also been discussed. Approaches for the
interpretation of kinetic data and modeling have been reviewed.
The type of reactor to be employed for carrying out ion exchange resin catalyzed
reactions has been discussed. A new emerging area is the adoption of distillation column
reactors (e.g., methyl tert-butyl ether from isobutylene and methanol).
Some aspects of the deactivation and regeneration of these catalysts have been briefly
discussed.
The use of ion exchange resins for the removal of impurities from mixtures and the
separation of close-boiling compounds by exploiting their different chemical reactivity has
been highlighted.
Suggestions have been made for further work.
Keywords: catalysis; cation exchange resin; alkylation; etherification; esterification; Am-

berlyst-15

Contents
Abstract
1. Introduction
2. Previous reviews
3. Preparation of ion exchange resin catalysts
4. Resin structure and its catalytic consequences

* Corresponding author.
0923-1137/93/$06.00 1993 - Elsevier Science Publishers B.V. All rights reserved

1
2
3
3
4

5.
6.
7.
8.

A. Chakrabarti, M.M. Sharma/React. Polym. 20 (1993) 1-45

Catalysis by ion exchange resins


Advantages and limitations of ion exchange resins as catalyst
Selectivityof ion exchange resin catalysts
Kinetics of ion exchange resin catalysis
8.1. Role of mass transfer
8.2. Kinetic model
8.3. Distribution coefficient
9. Reactions catalyzed by cation exchange resins: Reactions of industrial importance
10. Reactors used for ion exchange resin catalyzed reactions
11. Some important observations on ion exchange resin catalysis
12. Separation of close-boilingcompounds through reactions catalyzed by ion exchange resin
13. Removal of impurities from a mixture by reaction catalyzed by ion exchange resin
14. Modified ion exchange resins and supported perfluorinated resin sulfonic acid
15. Deactivation and regeneration of ion exchange resin catalysts
16. Ion exchange resins versus clays versus other catalysts
17. Conclusions
18. Future scope
References

1. Introduction
An ion exchange material may be broadly
defined as an insoluble matrix containing
labile ions capable of exchanging with ions in
the surrounding medium without major physical change taking place in its structure. The
resins typically employed are sulfonic acid
cation exchangers in the hydrogen form or
quaternary a m m o n i u m anion exchangers in
the hydroxide form. The idea that ion exchangers can be used as a substitute for
mineral acids and bases, dates back to 1911
[1]. The first commercial process, for esterification and ester hydrolysis over Wofatit phenolsulphonic acid-formaldehyde condensation polymers, was developed by I.G. Farbenindustrie in the 1940s in the context of
the German war effort [2]. Although ion
exchange catalysis involves the use of ion
exchange resins to promote reactions that
are normally catalyzed by mineral acids and
bases, there are many instances where the
use of ion exchange resin catalysts has led to
improved activity a n d / o r yield over the homogeneous catalysts. For the reaction of linear ketones and alcohols to give ketals, the
rates were impracticably low with the soluble

5
6
7
8
8
9
11
12
22
23
25
26
27
29
30
31
32
32

acids, but the use of acid ion exchange resin


provided reasonably high reaction rates at
the low temperatures employed [3]. In the
case of dehydration of tert-butanol, sulfonic
acid resin catalysts were found to be much
more effective than the soluble acid catalysts
[4]. Similarly, activities higher than soluble
p-toluenesulfonic acid ( p - T S A ) h a s been reported for the addition of alcohols to olefins
using ion exchange resin catalysts [5]. For the
etherification of ethylene and propylene glycols with isobutylene, it was observed that
the rates of reaction with cation exchange
resin catalyst were much higher than those
with soluble p-TSA catalyst [6]. In the case
of etherification of the phenols with isobutylene, the yields of the corresponding tertbutyl ethers were remarkably high compared
to that with p-TSA [7a]. Because of the
improved selectivity of the ion exchange
resins, the current ion exchange resin catalyst capacity of about 4000 m 3 / y r of resin is
expected to double over the next five years,
especially due to their high consumption in
methyl tert-butyl ether (MTBE) plants [7b].
This paper takes into consideration all the
previous reviews which have appeared in the
literature. In this review, emphasis has been

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45

especially given to the literature of the last


seven years. An attempt has been made to
focus attention on important aspects of catalysis by ion exchange resins. The new developments in preparing tailor-made resin catalysts have also been discussed. The use of
cation exchange resins for a variety of acidcatalyzed reactions has been covered; special
attention has been given to industrially important reactions. The potential of expensive
perfluorinated sulfonic acid based ion exchange resin as supported catalysts has been
considered.

2. Previous reviews
Bhagade and Nageshwar have published
reviews on different reactions, namely, hydration, dehydration, esterification, ester hydrolysis and alkylation, catalyzed by ion exchange resins [8-12]. In all the cases, they
have tabulated the reactants and the catalyst
used. Widdecke [13], in an exhaustive review,
has discussed the working concepts of ion
exchange resins as catalyst. Some important
points about the design and industrial applications of the catalysts have also been highlighted by him. Recently, Neier [14] has published a comprehensive review on the ion
exchangers as catalyst. Olah et al. [15] have
published a review article on perfluorinated
sulfonic acid resin catalyzed reactions. The
present review covers some additional and
new aspects of catalysis and also highlights
many of the industrially important reactions.
3. Preparation of ion exchange resin catalysts
Most present day ion exchange resins are
addition copolymers prepared from vinyl
monomers. These polymers have a higher
chemical and thermal stability than their
forerunners, the condensation polymers,
Also, addition polymerization has the advan-

tage that the degree of crosslinking and the


particle size of the resins are more readily
adjusted [16].
The most important resins of this type are
crosslinked polystyrenes with sulfonic acid
groups which have been introduced after
polymerization by treatment with concentrated sulfuric acid or chlorosulfonic acid
[17-19]. As a rule, divinylbenzene is used as
the crosslinking agent (commercial mixture
of divinylbenzene and ethylstyrene is used
for this purpose).
The copolymer beads are prepared by the
so-called pearl polymerization technique
[20,21]. The monomers (from which polymerization inhibitors have been removed) are
mixed and a polymerization catalyst such as
benzoyl peroxide is added. The mixture is
then added to a thoroughly agitated aqueous
solution which is kept at the temperature
required for polymerization (usually 85 to
100C). The mixture forms small droplets
which remain suspended. A suspension stabilizer (gelatin, polyvinyl alcohol, sodium
oleate, sodium methacrylate, magnesium silicate, etc.) in the aqueous phase prevents
agglomeration of the droplets. The size of
the droplets depends chiefly on the nature of
the suspension stabilizer, viscosity of the solution, and agitation and can be varied within
wide limits. The polymer is obtained in the
form of fairly uniform spherical beads. For
most purposes, a bead size of 0.4 to 1.1 mm
is preferred, but beads from 1 ~ m to 2 mm
in diameter can be prepared without major
difficulties. Polymerization in aqueous suspension is possible if the monomers are insoluble in water. This is not always the case.
However, water-soluble monomers such as
acrylic acid can be made insoluble, for exampie, by esterification. The ester groups are
then hydrolyzed after polymerization. Alternatively, the "water-in-oil" technique (inverse suspension polymerization) can be
used. Here, the copolymer is formed exclusively from hydrophilic monomers. The

droplets of the polymerization mixture are


suspended in an inert organic solvent in
which the monomers are insoluble,
The polymerization procedure described
above is generally adopted for manufacturing
gel-type resins. However, the conventional
pearl polymerization technique can be suitably manipulated to get highly porous, the
so-called macroreticular or macroporous, ion
exchange resins. An organic solvent which is
a good solvent for the monomers, but a poor
solvent for the polymer, is added to the polymerization mixture. As polymerization progresses, the solvent is squeezed out by the
growing copolymer regions. In this Way, one
can obtain spherical beads with wide pores
(several hundred Angstrom units) which allow access to the interior of the resin even
when non-polar solvents are used [22].
The sulfonation of the resins presents no
problems if proper precautions are taken,
Sulfonation progresses from the outer shell
towards the centre of the particle and is
accompanied by considerable swelling and
evolution of heat. The matrix is strained
rather severely, and the beads may crack,
Such disintegration can be avoided by letting
the beads swell, prior to sulfonation, in solvents such as toluene, nitrobenzene, methylene chloride or trichloroethane [19]. Similar
strains develop when the sulfonated beads
are transferred into a dilute aqueous solution
or pure water. Therefore, it is advisable to
transfer the beads at first into a highly concentrated electrolyte solution which causes
less swelling, and then to dilute the solution
progressively [19].
The commercial resins available are heterodisperse and have properties which can
be attributed directly to the band width,
within which the bead sizes fall. These irregularities can be observed in the mechanical
stability, hydraulic properties, and kinetic
phenomena in columns [23].
In the past few years it has been possible
to disperse liquid monomers into droplets of

A. Chakrabarti, M.M. Sharma/React. Polym. 20 (1993) 1-45

uniform size. If suitable methods are used to


prevent them from coalescing to form larger
droplets, monodisperse beads are formed,
yielding monodisperse resins after functionalization. Using the process developed by
Bayer AG, Leverkussen, it is possible to produce both gel and macroporous resins. For
monodisperse resin there is no Gaussian frequency distribution. Thus, "monodisperse"
refers to products where at least 90% of all
beads are of the same size with a tolerance
of + 0.1 mm. However, the beads are usually
within an even narrower tolerance, e.g. 90%
within +0.05 mm diameter.
Another type of ion exchange resin,
namely, perfluorinated sulfonic acid resin
(Nation), is prepared from a copolymer of
tetrafluoroethene and perfluoro[2-(fluorosulfonylethoxy)]polyvinyl ether. The perfluorinated vinyl ether is produced by reacting
tetrafluoroethene with SO 3 to form a cyclic
sulfone which subsequently rearranges to fluorocarbonyl methane sulfonyl fluoride. The
linear analog reacts with two moles of hexafluoropropylene oxide to yield a compound
with a terminal 1-fluorocarbonyltrifluoroethoxy group. This group loses carbonyl fluoride on heating with Na2CO 3 to give the
perfluorinated vinyl ether. The copolymer
resin in the sulfonyl fluoride form is base
hydrolyzed to the alkali form and then acidifled to the sulfonic acid form [24].

4. Resin structure and its catalytic consequenees


The resins can be divided into two groups
having major structural differences, gel and
"macroreticular" resins. On a microscopic
scale, the three-dimensional polymeric gel
matrix is homogeneous with no discontinuities. If the gel beads are totally dry, then the
polymeric matrix collapses and the polystyrene chains will be as close as atomic forces
allow. In this condition, unless a reactant is

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45

itself capable of swelling the matrix, the collapsed gel bead exhibit almost no catalytic
activity because only the sites on the bead's
surface (which are insignificant in number
compared to the total active sites within the
body of the bead) are available to the reactants [25]. Therefore, the swelling ability of
the reactants is a prerequisite for catalysis by
gel resins. Gel resins are characterized by
divinylbenzene content generally below 12%.
The limitation of the gel resins was largely
overcome, around 1960, with the advent of
the "macroporous" ion exchange resins [2630]. They consist of agglomerates of very
small microspheres interspersed with macropores. Reactants may move easily into the
interior of the bead through these macropores. The significant consequence is that,
whereas gel resins can only function effectively as catalysts in a swelling medium, macroreticular resins can function also in nonswelling solvents, thereby greatly expanding
the possibilities of resin catalysis [31-34].

5. Catalysis by ion exchange resins


The principal differences between catalysis by dissolved electrolytes and by resins are:
(i) with resins as catalysts, the catalysis
overlaps with diffusion, adsorption and desorption processes [35]; and the concentration of the reactants inside the catalyst (where
the reaction takes place) may be different
from that in the bulk solution, depending on
the nature of the matrix, fixed ionic group,
the reactants, and the solvents. For a particular catalyst and reactant system, the distribution of the reactants between the resin phase
and the bulk solution may be changed by
changing the solvent [36].
(ii) the matrix with the fixed ionic group
may have some influence (which is not purely
physical) on the course of the reaction,
In homogeneous catalysis, the catalytically
active species is dissolved in the reaction

medium and present uniformly throughout


the system. This is not the case with resin
catalysis. The catalytically active groups are
anchored to the matrix and, in the solventresin system, are located at the surface of
and within the body of the resin bead only.
The bulk solution is completely free of catalyst and consequently catalytically inert.
However, in polar solvents, acid leaching may
be observed leading to the presence of catalytically active species in the bulk solution
as well as in the resin.
Ion exchangers are, in effect, particulate
active species and when used as catalysts
offer the chemical benefits of homogeneous
catalysis combined with the physical and mechanical benefits of heterogeneous catalysts
[25].
Depending on the nature of the reactants
and solvent, all the resin-catalyzed reactions
fall into one of two major types, namely,
Type A and Type B [25]. Type A can be
subdivided into two groups: (A1) catalysis by
fully water-swollen resins in aqueous systems
and (A2) catalysis in mixed water/organo
solvents. In the first case, the hydrated proton is the catalytic agent and the resin is
found to be more effective than homogeneous acid. Davis and Thomas [37] showed
that the resin-catalyzed hydrolysis of a series
of esters using pure water as a solvent was
more effective than hydrochloric acid.
Haskell and Hammett [38] used 70% acetone
as the solvent in the hydrolysis of different
esters catalyzed by ion exchange resin as the
catalyst (Type A2) and found that hydrochloric acid was more efficient than the resin.
Here the presence of acetone changes the
distribution of the ester in favour of the bulk
solution which is rich in acetone (ion exchanger is enriched by water due to its higher
dielectric constant) leading to a lower efficiency of the resin. The above discussion also
shows the importance of reaction conditions
in achieving success in resin catalysis. Type B
is divided into two subgroups: (B1) catalysis

A. Chakrabarti, M.M. Sharma/React. Polym. 20 (1993) 1-45

by (water-free) resins in non-aqueous systems where water is not a product of the


reaction and (B2) non-aqueous catalysis in
which water is a by-product of reaction,
Alkylation of phenol with olefins is a typical
example of the Bl-type reaction. Water has a
detrimental effect on the Bl-type reactions
(which are catalyzed by undissociated sulfonic acid groups). The cause for the detrimental effect of water on Bl-type reaction
was given by Zundel [39]. Water competes
with the reactant for the active sites and is
strongly bonded to sulfonic acid groups. One
water molecule attaches itself to four sulfonic acid groups leading to a reduction in
the number of available acid groups which in
turn decreases the rate of the reaction. Examples of type B2 reactions are the dehydration of an alcohol, the esterification of acid
with alcohol, etc. Here as the reaction proceeds, transition from B2 to A1 takes place
with resulting drop in the catalytic activity,
Therefore to maintain the B-type character,
i.e., to maintain high catalytic activity, water
needs to be removed from the system as soon
as it is generated,

6. Advantages and limitations


change resins as catalyst

of ion ex-

The particulate nature of the ion exchange


resins permits mechanical separation of the
catalyst and the reactant-product mixture by
filtration or decantation, thus eliminating
distillation or extraction procedures for
product isolation. Resins can be used as catalyst in continuous processes in fixed bed reactors. In this case, size of the ion exchange
resin may necessitate the use of special support trays [40]. Particles should preferably be
of uniform size. Otherwise, smaller particles
may fill up the space between larger particles
causing improper distribution of the reactants in the resin bed [41].
Catalysis by ion exchange resins lowers

capital and processing costs by eliminating


the steps and equipment required for catalyst removal in analogous homogeneous processes. The waste disposal problem is obviated. Product purity and yield are improved
in many cases. Resins can often be used, in
hundreds of catalytic cycles without regeneration in well-designed processes. Therefore,
the cost of the resin is spread over the life of
the resin, making the catalyst cost per unit
production lower than their homogeneous
counterpart in real terms.
The ability of ion exchange resins to work
in both aqueous and non-aqueous systems
confers unique advantages. Hydrophobic systems can be catalyzed by dry macroreticular
resins. The drying of the resins is much easier than that of p-TSA. Resins even having
the equivalent strength of strong mineral acid
can be handled without danger to personnel.
Due to its heterogeneous nature the resin,
though it contains acid at high concentration,
can be used conveniently in mild steel equipment: the number of acid groups at the surface of the bead in contact with equipment is
a very small percentage of the total number
of acid groups present, thus obviating the
problem of corrosion. Agitation can further
reduce any corrosion by minimizing the contact between the resin particles and the
equipment.
The major disadvantage of ion exchange
resins in catalysis is their relatively low thermal stability. Ion exchange resins can withstand temperatures below 125C for long periods. For macroreticular resins, operation at
150C for a long time may cause desulfonation leading to a release of sulfonic acid as
well as a drop in the activity. It is recommended that such catalysts are not used
above 120C. Research is going on to develop ion exchange resins having higher thermal stability, and this will be discussed later
in the text. Nation can withstand temperature up to 200C, but unfortunately such
materials are expensive; the use of supported

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45

perfluorinated sulfonic acid may however reduce overall operating costs,

7. Selectivity of ion exchange resin catalysts


The selectivity of ion exchanger catalysts
often leads to yields which are considerably
higher than those obtained with catalysis by
dissolved electrolytes [16]. The ion exchanger
distinguishes more sharply between the various reactant molecules. Ion exchangers as
catalysts are perhaps about halfway between
the non-selective dissolved electrolytes and
the extremely selective enzymes. Ion exchange resins can be tailor-made to use as a
means for selective catalysis of small
molecules in the presence of large molecules.
A crosslink density may be chosen to inhibit
matrix expansion sufficiently so that large
molecules are excluded from the interior of
the bead by sieve action while small
molecules may enter the matrix thus favouring the reaction of the small molecules only
[25]. Other specific examples of selective resin
catalysis, which cannot be explained simply
in terms of molecular size and sieve action,
include condensation of furfural with an
aldehyde. Such effects may be tentatively
explained by differences in sorption of various reactants. The preferential sorption of
one of the reactants in the resin may favour
a reaction which is insignificant in catalysis
by dissolved electrolytes. In the above-mentioned case, dissolved bases under otherwise
identical conditions almost exclusively give
self-condensation of the aldehyde,
Ion exchange resins showed higher selectivity for the desired product in many reactions. In the alkylation of benzene with
styrene, the percentage yield of alloylated
product was 0, 10.7 and 13.7 with p-TSA,
Amberlyst-15 and Nafion-H as the catalyst,
respectively [42]. In the alloylation of toluene
with styrene, the percentage yield of the
alkylated product was 51.5, 61.5 and 86.2

with p-TSA, Amberlyst-15 and Nafion-H as


the catalyst, respectively [42]. Etherification
of phenol with isoamylene was tried at low
temperature with p-TSA as the catalyst [7].
Only C-alkylated product was formed. However, etherification proceeded well at the
same temperature with Amberlyst-15 as the
catalyst.
In the alkylation of phenol with olefins,
some remarkable differences were observed
with ion exchange resin as the catalyst in the
o-/p-product distribution [74]. With a-methylstyrene (AMS) as the alkylating agent, the
o-/p-distribution was 16:84 and 3:97 with
p-TSA and Amberlyst-15 as the catalyst, respectively; with diisobutylene (DIB) as the
alkylating agent, the values were 41:59 and
5 95 with p-TSA and Amberlyst-15 as the
catalyst, respectively; and with 1-octene as
the alkylating agent, the ratio was 72 28 and
59:41 with p-TSA and Amberlyst-15 as the
catalyst, respectively. Recently, Chandra and
Sharma [44] have observed that methyl tertbutyl ether (MTBE) can be advantageously
used for the butylation of phenol with Amberlyst-15 as catalyst. The selectivity for 2tert-butylphenol can be improved by carefully
selecting the temperature and the level of
conversion. For the alkylation of substituted
hydroquinones such as methylhydroquinone,
phenylhydroquinone and tert-butylhydroquinone, Kim et al. [177] have reported better selectivity to 2,5-di-tert-butylhydroquinones with Amberlyst-15 catalyst than 70%
sulphuric acid. The selectivity of 2,5-substituted hydroquinones was in the descending order of tert-butyl, phenyl, and methylhydroquinones.
In the acid-catalyzed cracking of MTBE,
considerable improvement in the selectivity
for the desired product could be realized by
using a surface-sulphonated porous polystyrene catalyst instead of the commercially
available resins. This may be attributed to
the formation of by-products in the solid
matrix regions of the commercial catalysts

8
which is inhibited by providing the -SO3H
groups only on the surface of the catalyst
[45].

8. Kinetics of ion exchange resin catalysis


8.1. Role of mass transfer
The theory of catalysis by ion exchangers
has not yet been well-established. In all cases
which have so far been studied and in which
limitations due to slow diffusion process can
be ruled out, the kinetic order of the chemical reaction in the particles was found to be
the same as in homogeneous catalysis by a
dissolved electrolyte. This is strong evidence
in support of the view that the reaction
mechanism is essentially the same in both
cases. However, there are differences in reaction rates, even when the amount of catalyst (in counter-ion equivalents) is the same
in both cases. A significant difference between homogeneous catalysis and heterogenized homogeneous catalysis is the presence
of diffusional processes in the later case.
Most of the ion exchange resin catalyzed
reactions reported in the literature have been
carried out in the liquid phase; therefore,
our discussion will be limited to the liquidphase reactions. However, the same treatment can be adapted, with a few obvious
changes, to gas-phase reactions. In order for
the reaction to occur, reactant molecules
must first migrate from the bulk solution into
the ion exchanger; and must react; the reaction products must then migrate back into
the solution. Thus, three phenomena can
affect the rate of overall process: (i) diffusion of the reactants and products across the
film (Nernst diffusion layer) surrounding the
catalyst particle; (ii) diffusion of the reactants and products in the interior of the
catalyst particle; and (iii) chemical reaction
at the active sites. If film diffusion is much
slower than the chemical reaction, it must

A. Chakrabarti, M.M. Sharma/React. Polym. 20 (1993) 1-45


obviously be rate-controlling since in this limiting case all reactant molecules react as
soon as they reach the surface of the catalyst
particle. If the chemical reaction is much
slower than the diffusion processes, sorption
equilibrium of the reactants is established
and upheld throughout the catalyst particle
since diffusion is enough to make up for the
disappearance of reactants by chemical reaction. The overall rate is thus controlled by
the rate of the chemical reaction throughout
the particle. If intraparticle diffusion is much
slower than the actual chemical reaction, the
reactant molecules will react before they have
time to penetrate into the interior of the
catalyst particle. In this latter limiting case,
the reaction occurs only in a thin layer at the
surface and its rate is controlled by either
film diffusion or chemical reaction at the
surface. If the film diffusion is sufficiently
fast, the process is controlled by chemical
reaction at the surface. Intraparticle diffusion can affect the overall rate of reaction,
but it can never be the sole rate-determining
step as diffusion and chemical reaction
are parallel steps, occurring simultaneously
within the resin bead.
These considerations suggest that, in the
limiting cases, the overall rate of any ion
exchange resin catalyzed reaction is controlled by: (i) film diffusion; (ii) chemical
reaction throughout the particle; and (iii)
chemical reaction at the surface. Obviously
intermediate cases arise if the rates of individual steps are comparable. However, it is
found generally that film diffusion does not
control the overall rate in ion exchange resin
catalyzed processes. If the viscosity of the
reactant mixture is too high or the speed of
agitation is too low, then film diffusion may
control [16]. But these are exceptional cases.
Often the overall rate is controlled by a
combination of intraparticle reaction and intraparticle diffusion.
The procedure for discerning the controlling mechanism in ion exchange resin cat-

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45


alyzed liquid/solid reaction has been discussed in detail by Doraiswamy and Sharma
[46]. They have also discussed the procedures
for discerning the controlling mechanism in
gas/liquid/solid (where the solid acts as the
catalyst) reactions carried out in slurry or
fixed bed reactors [47]. Carra [48] has discussed the type of reactors used, role of
hydrodynamic factors, mass transfer in agitated systems, role of intraparticle diffusion,
etc., for ion exchange resin catalyzed reactions,
8.2. Kinetic model
The scope of catalytic application of
strongly acidic ion exchange resins in the
liquid phase extends from essentially homogeneous mechanisms to classical heterogeneous catalysis. Depending on the interaction of the solvent with the acid groups and
the polymer backbone, different degrees and
possibly types of swelling occur. The idealized homogeneous state requires complete
swelling and total dissociation of the polym e t - b o u n d - S O 3 H group (specific catalysis),
Whereas, the heterogeneous state is characterized by a direct interaction of the substrate with the polymer-bound -SO3H group
(general catalysis). Most resin-catalyzed reactions can be classified either as quasi-homogeneous (e.g., hydrolysis of sucrose, carboxylic esters) or quasi-heterogeneous (e.g.,
benzene propylation, isobutylene oligomerization). Complicated situations arise where
the solvent causes only partial dissociation of
-SO3H groups, due to a low degree of solvation (e.g., MTBE synthesis) [49].
The rate of a resin-catalyzed reaction will
be a function of (1) the acidity and (2) the
accessibility of the active sites [13], i.e.,
r = f(acidity, accessibility)
where
acidity = f(type and number of acid groups,
divinylbenzene content, reaction medium)
and

9
accessibility = f(divinylbenzene content, particle size, porosity, reaction medium, diffusivity).
For a given type of acid resin catalyst,
higher divinylbenzene content will reduce the
swelling ability of the resin, thus reducing
accessibility to the catalyst. In the presence
of more polar reaction medium, the
polymer-bound-SO3H groups are more dissociated leading to a decrease in the acidity
of the sulfonic acid resin. With more polar
components, the swelling of the resin is
higher and, hence, the accessibility is improved. Buttersack [49] has further classified
the accessibility into non-interaction and interaction accessibility. The former is akin to
the term 'dispersion' used in the conventional heterogeneous catalysis for defining
the fraction of catalytically active groups at
the surface and is related to the morphology
of the resin in the dry state. The latter reflects the influence of the solvent and adsorbed molecules (the steady state reaction
medium) on the ion exchange resin catalyst.
The accessibility is also a function of the
effective diffusivity of the reactants within
the pores, which is related to the Thiele
modulus through the equation
$=Rfk/D
where R is the radius of the catalyst particle,
k is the reaction rate constant, and D is the
effective diffusivity of the reactant.
Apart from the adsorption isotherm and
distribution coefficient of the reactant, the
Thiele modulus also affects the actual number of sites, out of the total accessible sites,
occupied by the reactant. The effect of intraparticle diffusion can be incorporated in the
rate expression by way of the effectiveness
factor '~' which can be correlated to the
Thiele modulus. A two-phase model has been
suggested for the effectiveness factor of a
macroreticular resin [50], since it is essential
to take account of both macropore and micro-phase diffusion.

10
In the absence of any intraparticle diffusion limitation, the rate of reaction can be
expressed using a simple pseudohomogeneous model or a more complex model based
on the Langmuir-Hinshelwood mechanism,
The pseudohomogeneous model is based on
the Helfferich approach which treats the
catalysis by ion exchange resins as homogeneous catalysis confined within the internal
catalyst mass (i.e., resin volume), wherein
the reactants, products and solvents are in
distribution equilibrium with the bulk solution. Here, the reaction mechanism is similar
to the one observed with the dissolved electrolytes. The difference in the performance
of the two is attributed to the selective sorption of one of the reactants by the resin
depending upon the nature (e.g., polarity)of
the reactants/solvent medium. In this model,
linear relationships are assumed between the
reactant concentration at the active sites
within the resin and that in the bulk. Buttersack [49] has defined a catalytic efficiency (q)
as the ratio of the overall reaction rate per
acid site, obtained with a resin catalyst to
that obtained with a mineral acid. The q
values for many reactions have been found to
be identical to the distribution coefficients,
The pseudohomogeneous approach is particularly applicable in cases where one of the
reactants or solvent is highly polar. In the
presence of polar components, solvation of
the polymer-bound -SO3H group takes
place, and the catalysis is by the solvated
proton. For example, this model has been
successfully used for the hydrolysis of sucrose [51], the hydrolysis of carboxylic esters
[51], the hydration of olefins [52,53], and the
dehydration of alcohols [54].
Where, due to the non-polar nature of the
reactant or solvent, the polymer-bound
undissociated -SO3H group is the proton
donor, a simple homogeneous model may
not fit the data. A more suitable approach is
one based on the Langmuir-Hinshelwood or
Rideal-Eley mechanism. These have been

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45

satisfactorily applied in the case of isobutylene oligomerization [55], dimerization of


isoamylene [56], and benzene propylation
[34]. A more complex situation arises for the
etherification reactions of olefins with alcohols ( M T B E / E T B E / T A M E synthesis): depending upon the concentration of the alcohol, there can be a transition from general to
specific catalysis. For MTBE synthesis, at
lower concentration of methanol the catalysis
is by the polymer-bound -SO3H groups
which are very active but if the concentration
of methanol is increased, methanol dissociates the acid groups and solvated protons
become the catalytic agents which are relatively less active. At very high methanol concentrations, the methanol is no longer able
to affect the rate either as reactant or as
solvent, having completely levelled the catalyst activity. Some researchers (Voloch et al.
[57], Subramaniam and Bhatia [58]) have used
a pseudohomogeneous model in the case of
MTBE synthesis and have been able to correlate the experimental kinetic data satisfactorily; nonetheless Langmuir-Hinshelwood
or Rideal-Eley models have been more
widely used (Hoffmann et al. [59-62], Ancillotti et al. [5], Gicquel and Torck [63], Gates
and Johnson [64], Cunill et al. [65]), and
these can be applied and extrapolated over a
wider range of concentrations. When the
rate-limiting step involves a reactant from a
fluid phase with one that is adsorbed, the
mechanism is called a Rideal-Eley mechanism. If all the reactants involved in the
rate-determining step are adsorbed, the corresponding kinetic behaviour is termed
Langmuir-Hinshelwood. A few reactions exhibit maximum of reactivity at certain substrate concentration (e.g., vapour-phase dehydration ofalcohols, vapour-phasesynthesis
of M T B E / E T B E [66]). Such a maximum can
be explained by a Langmuir-Hinshelwood
mechanism, based on two-site adsorption of
isobutene and the ether, and a simple adsorption of methanol. Buttersack [49] has

11

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45

suggested a more general concept of variable


active centres, in which, at low degree of
adsorption, each substrate molecule requires
a certain number of acid sites to form an
active centre. As is known for a usual Langmuir isotherm, a fraction of sites participate
in the adsorption and another fraction of
sites remain unoccupied. When all sites participate in the adsorption, the maximum reaction rate is achieved. After attainment of
saturation, further adsorption of the substrate molecules is possible by lowering the
size of the active centres down to one acid
group per substrate molecule. This size-reduction of the active centres causes a decrease of the acidity and the reaction rate.
Thus the maximum of reactivity depends on
the proton acceptor ability of the substrate;
the higher the ability the sharper is the maxim u m and the lower is the substrate concentration at maximum reactivity. In the extreme case when the proton acceptor ability
of the substrate is very low, the maximum
disappears at infinite substrate concentrations, and the general kinetics reduces to the
Rideal-Eley kinetics. When operating in a
solvent with higher proton acceptor ability
than that of the substrate, the additional
influence of the substrate can be neglected,
and the status of the acid sites is nearly
exclusively determined by solvation, so that
an active centre should be identical with a
single solvated acid site. Thus, the deactivation of the resin catalyst by the substrate has
been explained satisfactorily by the concept
of variable active centres for the propylation
of benzene (Buttersack et al. [67]).

Reactions under supercritical conditions


The manufacture of isopropanol by hydration of propylene and sec-butanol by hydration of butenes may involve supercritical conditions. Here fixed bed reactors are employed. There is hardly any literature on the
kinetics in such reaction systems [68].

8.3. Distribution coefficient

As discussed earlier, when catalysis by the


cation exchanger is similar to homogeneous
catalysis, the actual concentration of the different components of the reaction mixture in
the resin phase may not be the same as that
in the bulk solution, since the microenvironment of the resin may differ substantially
from the bulk. There may be preferential
distribution of some of the components depending on their relative compatibility with
the resin phase. Therefore, knowledge of the
distribution coefficients is necessary to find
out the actual concentration of the different
components in the resin phase. Distribution
coefficient of a particular component is defined as the ratio of the concentration of that
component in the resin phase to that in the
bulk phase. Distribution coefficient can be
found out by the method described by Helfferich [69,70]. Two methods have been suggested for this purpose. In the first one, the
reactants are equilibrated with the catalyst
from which the catalytically active species is
replaced with some inactive ions. It may not
present the actual picture; the replacement
of catalytically active species by some inactive ions must change the environment within
the resin phase, leading to a change in the
distribution of the components. The other
method, which is more accurate, consists of
equilibrating small particles of the catalyst
with the reactants to ensure the attainment
of sorption equilibrium before any appreciable conversion occurs. Then the contents of
the resin are eluted by some solvent, and the
reaction is completed in the eluate by some
soluble catalyst. The amount of the products
in the eluate (which is stoichiometrically
equivalent to the amount of reactant initially
taken up by the resin) is then analytically
determined. In this method, there is no provision for estimating the distribution coefficient of the product which needs to be considered in the case of a reversible reaction.

12

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45

TABLE 1
Physical properties of some cation exchange resins
Catalyst

Amberlyst-15
Amberlyst-35
Amberlyst XN-1010
Nation NR 50
Lewatit SPC 118

Properties
Crosslinking density
(% DVB)

Surfacearea
(m2/g)

Pore volume
(ml/g)

Capacity Maximumoperating
(meq/g)
temperature (C)

20
20
> 50
_a
18

45
44
540
0.003
40

0.35
0.41
0.47
non-porous
0.46

4.8
5.2
5.3
0.8
5.0

120
140
150
200
120

a Not applicable.

A n o t h e r m e t h o d (adopted in our laboratory)


which may be more useful is as follows. The
reaction is carried out for some time. The
resin phase is then carefully filtered off from
the bulk solution. The contents of the resin
phase are then eluted by some suitable solvent and analyzed by the usual method. The
bulk phase is then analyzed as usual. Thus by
knowing the composition of the resin phase
and the bulk phase, the distribution coefficient of the reactants and the products can
be found out.

9. Reactions catalyzed by cation exchange

resins: Reactions of industrial importance


All the reactions generally catalyzed by
mineral acids, can be carried out in the presence of cation exchange resin as catalyst with
some of the advantages which have been
discussed earlier. The physical properties of
relevant cation exchange resins are listed in
Table 1. A n u m b e r of cation exchange resin
catalyzed processes have been commercialized, a list is given in Table 2. Table 3 lists

TABLE 2
Some important commercial processes catalyzed by cation exchange resins
Name of the product

Key reactants

Ref.

Methyl tert-butyl ether (MTBE)


Ethyl tert-butyl ether (ETBE)
tert-Amyl methyl ether (TAME)
Isobutylene
tert-Butanol
Isopropanol
sec-Butanol
Isopropyl esters
sec-Butyl acetate
sec-Butyl acrylate
Acrylates and Methacrylates
Alkylphenols
Bisphenol A
Methyl glucoside
Methyl isopropenyl ketone
Methyl isobutyl ketone

Methanol + Isobutylene
Ethanol + Isobutylene
Methanol + Isoamylene
tert-Butanol
Isobutylene + Water
Propylene + Water
n-Butenes + Water
Propylene + Carboxylicacids up to C20
Butene + Acetic acid
1/2-Butene + Acrylic acid
Acrylic acid or Methacrylic acid + Alcohols
Phenols + Olefins
Phenol + Acetone
Dextrose + Methanol
Methyl ethyl ketone + Formalin
Acetone + H 2

77,133
77
77
134-137
137
138-140
141, 142
143
144
145
146-148
149
149
150
151
152

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45

13

TABLE 3
Examples of the reactions catalyzed by cation exchange resins
Reaction

System

Catalyst

Ref.

Glyoxal + 3-Hydroxypropene
Formaldehyde + Methanol
Furfural + Ethylene glycol
Benzophenone + Trimethyl orthoformate
Acetone
Butanone
~ + Methanol
Cyclohexanone]
[Ethylene glycol
Cyciohexanone + ~ 1,3-Propanediol

Sulfonic acid resin


Lewatit SPC 118
D61, D72
Nafion-H

153
104
154
155

Dowex 50

Amberlyst-15

156

Nation-H
Amberlyst-15,
Dowex 50W-X2,
Dowex 50W-X4,
Dowex 50W-X12
Amberlyst- 15,
Amberlyst XN-1010
Nafion-H
Nation-H,
Amberlyst-15

157

Acetal formation

Alkylation of benzene / substituted benzene


Benzene + Ethylene
Benzene + Propylene

Benzene + 1-Dodecene
m-Xylene + Isoprene
Benzene + Styrene

34
158
159

idem.

42
42
42
42
42
42
42
42
42

Amberlyst-15

160

Phenol + Ethylene
Phenol + Propylene

Nafion-H
Nation-H,
Amberlyst-15

161

Phenol + Isobutylene (IB)


Phenol + Propylene
Phenol + Butenes
Phenol + IB
Phenol + Isoamylene (IA)
Phenol + 1-Octene
Phenol + Diisobutylene (DIB)
Phenol + a-Methylstyrene
Phenol + a-Methylstyrene
Phenol + IB
Phenol + DIB

idem.

Toluene + Styrene
p-Xylene + Styrene
o-Xylene + Styrene
m-Xylene + Styrene
1,2,4-Trimethylbenzene + Styrene
1,3,5-Trimethylbenzene + Styrene
Toluene + p-Methylstyrene
m-Xylene + p-Methylstyrene

idem.
idem.
idem.
idem.
idem.
idem.
Nafion-H

Alkylation of naphthalene
Naphthalene + Propylene

Alkylation of phenol / substituted phenols

Amberlyst-15,
Amberlyst XN-010,
Nation NR-50

idem.
idem.
idem.
idem.
Amberlyst-15
Sulfonic acid resin
Strongly acidic ion
exchanger

161
161
43
43
43
43
43
43
43
163
162
164

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45

14
TABLE 3 (continued)
Reaction

System

Catalyst

Ref.

Alkylation of phenol / substituted phenols


Phenol + DIB

Amberlyst-15,
Amberlyst XN-1010

Phenol + 1-Dodecene
Phenol + 1-Decene
Phenol + C8_18 n-oletins
Phenol + Propylene trimer
Phenol + Cyclohexene

idem.

Phenol + Styrene
Pyrogallol + Styrene
Phenol + 1-Pentanol
o-Cresol + 4-Methyl-l-pentene
m-Cresol + IB
p-Cresol + IB
m-Cresol /
p-Cresol /
2,5-Xylenol~ + AMS
2,4-Xylenol /
2,6-Xylenol}
m-Cresol
p-Cresol ~ + DIB
2,6-Xylenol)
Phenol
/
o-Chlorophenol!
+ IB
o-Ethylphenol
Guaiacol
Phenol
o-Chlorophenol| _.
+ IA
o-Ethylphenol /
Guaiacol
)
Phenol + MTBE
m-Cresol + MTBE
Hydroquinones + IB

KU-2
Sulfo-based cation exchangers
Amberlyst XN-1010
Amberlyst-15,
Amberlite IR-120,
Nation NR-50
Nation 117

idem.
Wofatit KPS
Amberlyst-15
Nation-H,
Amberlyst-15
Amberlyst- 15

165
165
166
167
168

124
169
170
171
172
173
174

idem.

100

idem.

100

idem.

7a

idem.

7a

idem.
Wofatit OK-80
Amberlyst-15

44, 175
176
177

a-Naphthol ~
t-Naphthol~ + IB

idem.

7a

a-Naphthol~ _.
/3-Naphthol~ + IA

idem.

7a

idem.

178

Phenylsalicylate + IB

Wofatit OK-80

179

Formaldehyde acetal +
Carbon monoxide

Unspecified commercial
ion exchanger

180

Alkylation of naphthols

Alkylation of poly( p-vinylphenol)


Poly(p-vinylphenol) + tert-Butanol

Alkylation of phenylalkoxy- and hydroxy-benzoates


Carbonylation

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45

15

TABLE 3 (continued)
Reaction

System

Catalyst

Ref.

Methyl N-phenylcarbamate + Formaldehyde


Anethole + Acetaldehyde
Phenol + Acetone

Cation exchange resin

idem.

181
182

Condensation

Phenol + Acetone
o-Benzoylbenzoic acid

1,4-Bis(2-propanol)benzene + Aniline
Cyclohexanone
to-Undecylenic acid

Revachit SPC-118 BG
(20% neutralized
with H2NCH2CHzSH)
Cation exchange resin
Sulfonic acid resin
Perfluorinated
sulfonic acid resin
Amberlite IR-120,
Dowex 50
Amberlyst- 15

183
184-189
190
191
192
193

Cross-dimerization
ot-p-Dimethylstyrene + IA

idem.

194

Aniline + Acrylonitrile

Strongly acidic resin

195

Lewatit SPC 118,


Amberlyst-15
SiO2-modified
Lewatit SPC 118
Amberlyst XN-1010,
Amberlite IR-118
Vionit CS-1

65,
196

Indion-130
Wofatit OK-80
Amberlyst-15, Polystyrene
sulfonic acid
Sulfonated polystyrene resin
Macroporous cation
exchanger
Nafion-H,
Amberlyst-15
Nafion NR-50,
Amberlyst-15, K2661,
Lewasorb AC 10 FT
Ion exchange resin

199-201
202

209
210

Strongly acidic
ion exchange resin
Ion exchange resin

211
212

Lewatit SPC
KU-23

213, 214
215

Amberlyst-15

95

Cyanoethylation
Decomposition / cracking
MTBE
TAME
Cumene hydroperoxide

p-tert-Butyl(cumene-hydroperoxide)

197
127
198

Dimerization / oligomerization
IB
IB
IA
1-Butene, 2-Butene
DIB
Styrene
AMS

Formaldehyde

203
204
205
206-208

Epoxidation
Ricebran oil + Hydrogen peroxide
Cottonseed oil + Hydrogen peroxide

Esterification of olefins
Propene +Acetic acid
n-Butenes + Acetic acid
1-Butene
Isobutylene
1-Octene ] +Acetic acid

1-Dodecene]

16

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45

TABLE 3 (continued)
Reaction

System

Esterification of olefins
Cyclohexene + Acetic acid
Cyclohexene + Acetic acid
Cyclohexene + Acetic acid + water
/Formic acid
Camphene + ~Acetic acid
Dihydromyrcene + Acetic acid
Styrene + Acetic acid
Styrene +Acetic acid

Propylene + Chloroacetic acid


1-Butene + Acrylic acid
Camphene + Methacrylic acid
Dicylopentadiene + Acetic acid
2_Methyl_l_butene/
Isoprene
) + Acetic acid
Cyclopentadiene ]
Propylene + Myristic acid
Esterification of alcohols
Isopropanol + Acetic acid

Hexyl alcohol + Acetic anhydride


Glycerol +Acetic anhydride
fAcetic acid
Glycerol + ~Propionic acid
Ethanol + Oxalic acid
C 6-12-Alchls + Maleic anhydride
Methanol + Methacrylic acid
2-Ethylhexanol + Methacrylic acid
2-Ethylhexanol + Acrylic acid
Ethylene glycol
)
Di-ethylene glycol ~
Tri-ethylene glycol [ + Acrylic acid
Tetra-ethylene glycol]
Methanol + Fatty acid

Catalyst

Ref.

001 x 7 (Cation exchanger)


Amberlyst-15
Lewatit SPC 118

216
217
218

Cation exchange resin

219

idem.
Ostion KS
Amberlyst-15, Amberlyst
XN-1010, Nation NR-50,
Monodisperse K2661,
K2631
Amberlyst-15
Lewatit SPC 118
D001-ML
Lewatit SPC 108

219
220

Lewatit SPC
118 (45%
neutralized
with K + )
Amberlyst-15

221
95
145, 222
223
224

225
226

Diaion HPK 55
(Modified to get
corresponding
fluorinated ion
exchange resin)
Ion Exchanger
EX 146H

227
228
229

D61, D72, Amberlyst-15

230

Amberlite IR-120
Vionit
Cation exchange resin
idem.
Diaion PK 208

123, 231
232
233
234
235

Strongly acidic
ion exchanger

236

Lewatit

237

4-Oxo-pentanoic acid
r)
c/s-Pinanoic acid
trans-Aconitic acid
+ Methanol Amberlyst-15
(S)-Mandelic acid
Oleic acid
Malonic acid monoethyl este
Phenol + Methacrylic acid
idem.
Ring opened adduct of cyanoethyl alcohol
and e-Caprolactone + Acrylic acid
idem.

238

239
240

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45

17

TABLE 3 (continued)
Reaction

System

Esterification of alcohols
Lipids + Acetic anhydride
Pyrrole diacids + methanol
Ethylene glycol
monoethyl ether + Acetic acid
Dimethylphosphoramidothioate
+ Acetic anhydride
Trioxane + Carbon monoxide
+ methanol

Catalyst

Ref.

Cation exchange resin


Amberlite IR-120
H-form cation
exchanger

241
242

Amberlyst-15
Strongly acidic
ion exchanger

244

243

245

Etherification
IB + Methanol

Cation exchange resin

IB + Ethanol

Amberlyst- 15,
Amberlyst-35,
Fluorocarbon sulfonic
acid
Amberlyst-15
Vionit CS-34
Amberlyst-15,
Amberlyst-35
Acidic ion
exchange resin,
Amberlyst-15
Amberlyst-15

IA + Methanol
2-Methyl-l-butene + Methanol
IA + Ethanol
IB + Ethylene glycol

IB + Propylene glycol
Ethylene
/Glycol
1-Butene + ~Propylene
|Glycol
I,Di-ethylene glycol ether
Ethylene
|Glycol
Propene + ~Propylene
|Glycol
~Di-ethylene glycol ether
Dicylopentadiene + Ethylene glycol
[Ethanol
Camphene + ~Isopropanol
Dihydromyrcene + Methanol
DIB
Methylpentene + Methanol
Pinene
a-Pinene~
~-Pinene] + Methanol

5, 77,
246-261
77,
262,
263
77, 264-266
267
267

268-270
6, 271

idem.

270

idem.

270

idem.

272

Cation exchange resin

273

idem.

273

Amberlyst-15

274

idem.

275

Cyclohexene + 2-Methyl-2-phenyl ethanol


Cyclohexene + Phenethyl alcohol
f Isoamyl alcohol
Styrene + ~Cyclohexanol

idem.
idem.

276
277

idem.

278

AMS + Methanol

idem.

91

A. Chakrabarti, M.M. Sharma 'React. Polym. 20 (1993) 1-45

18
TABLE 3 (continued)
Reaction

System

Catalyst

Ref.

idem.

7a

idem.

7a

Nation 511,
Amberlyst-15

150,
279, 280

Wofatit OK 80

281

Cation exchange resin


Imac C8P
XE-307
Amberlyst-15,
Amberlyst XN-1010
Amberlyst- 15,
Amberlyst XN-1010,
Lewatit SPC 118,
Lewatit SC 102

282
52
53

Cation exchange resin

288

D001-ML, Strongly
acidic ion exchange resin
Amberlyst-15
Lewatit SPC 118,
Amberlite XE-307
Amberlyst-15,
Amberlite IR-120
Duolite C467,
Amberlite ICR-718

289,
290
291
88-90,
218

Diaion-PK228
Ion exchanger

294b

E therific a tion
(Phenol
IB + {a-Naphthol
~/3-Naphthol
(Phenol
IA + {a-Naphthol
~/3-Naphthol
Glucose + Methanol
|

Methylolurea +
2-Methylpropanol

Hydration
Propylene + Water
Propylene + Water
n-Butenes + Water
IB + Water
Isopentenes + Water

Camphene
a-Pinene
Dihydromyrcene + Water
Dipentene
Carene
Camphene + Water
Citronellol + Water
Cyclohexene + Water
AMS + Water
Acrolein + Water
Cyclohexene
oxide + Water
2-Cyclohexenone + Water

84, 87, 283

284-287

91
292,
293

containingSO3H
Acrylonitrile + Water

or NH4OH
Strongly acidic
ion exchanger
containing Cu ions

294b

Ion exchange resin


Nation NR 501
Amberlyst-15

295

Hydrolysis
sec-Butyl-trifluoro acetate + Water
Ethylene glycol mono-tert-butyl ether + Water
Ethylene glycol di-tert-butyl ether + Water

Methyl Acetate + Water

idem.

296
297
298
299

Benzophe.none
dimethyl acetal + Water

Nation-H

155

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45

19

TABLE 3 (continued)
Reaction

System

Catalyst

Ref.

2,2-Dimethyl- 1,3-dioxolane + Water


Di-methylphosphoenolpyruvate + Water

Amberlyst- 15
Dowex 50H

300
301

endo-Dicyclopentadiene
2-Methyl-l-butene
1-Butene
2,5-Dimethylhexadienes
Styrene oxide

Strong acid ion exchanger


Sulfonic acid resin
Cation exchange resin
Duolite C26H
Nation 501,
Amberlyst-15
Amberlite IR-120
Dowex MSC- 1
Dowex MSC- 1
containing
(HSCH 2CH 2)2 NH

302
303
304
305

Ambe rlyst- 15

310

Isobutylene + Formaldehyde
Isoeugenol + Paraformaldehyde
a-Methylstyrene /
Styrene
) + Paraformaldehyde

Cation exchange resin


idem.

311
312

Cation Exchanger

313

N-(Ethanol)benzamide
2,3-Dimethyl-2,3-butanediol
4-Hydroxy-acetophenone
oxime + Acetic acid + Acetic anhydride
Phenylbenzoate

Rexyn- 100
Amberlyst-15

314
315

idem.
Nation

316a
316b

Isobutylene + Acrylonitrile

Indion CXC-125

317

1,2-Ethanedithiol + Benzophenone

Nafion-H

155

Ethanedithiol + ~Benzaldehyde
2-Furfural

Amberlyst- 15

318

1,2,6-Trimethyl-4-tert-butyl-benzene + Toluene
1,2,6-Trimethyl-4-tert-butyl-benzene + Biphenyl
2-Amino-4-tert-butyl toluene + Toluene
2,6-Di-tert-butyl-p-cresol + Toluene

Nafion-H
idem.
idem.
idem.

319, 320

Ethylene carbonate + Methanol

Acidic ion
exchange resin
Amberlyst-15
idem.

Hydrolysis

Isomerization

Epoxides
2,4'-Dihydroxy-2,2'-diphenylpropane
2,4'-Dihydroxy-2,2'-diphenylpropane

Mixture of 3-methyltetrahydrophthalic
anhydride and 4-methyltetrahydrophthalic
an hydride

306
307
308

309

Prins reaction

Rearrangement

Ritter reaction
Thioacetalization

Transalkylation

Transesterification

Dodecyl acetate + n-Butanol


Cyclohexyl acetate + n-Butanol

321
95
217

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45

20
TABLE 3 (continued)
Reaction

System

Catalyst

Ref.

3-Methoxy-l-butene + Propylene glycol

Dowex MSC-1-H

322

Isophoronenitrile + Ammonia
1-imino-3-cyano-3,5,5trimethylcyclohexane
N-Hydroxymethylacrylamide +
Ethyl alcohol --* N-Ethoxymethylacrylamide
Aniline + Phthalic anhydride
---, N-Phenylphthalimide
4,4'-Diaminodiphenylmethane + Phthalic anhydride
4,4'-Diphthalimidodiphenylmethane
Maleic anhydride + 4,4'-Methylenedianiline ~ N,N'-(Methylenedi-p-phenylene)bismaleimide
1,3,5-Trioxane + Alkyl cyanide
1,3,5-Triacylperhydro-l,3,5-triazines
Acetone + Ammonia ~ 2,2,6,6-Tetramethyl4-piperidinone
1,1-Diethoxybutyraldehyde
--* 2,5-Diethoxytetrahydrofuran
Acetone + Glycerol ~ 2,2Dimethyl-4-(hydroxymethyl)-l,3-dioxolane
Glycidyl methacrylate +
Acetone ~ Dithiolane
Xanthan gum + Propylene
oxide --, Hydroxypropylated xanthan gum
2-Methyl-l,4-diacyloxy1,3-butadiene + Hydrogen
peroxide ~ 2-Methyl-4,4-diacyloxy-2-butenal
Coumarone resin raw oil +
Distillate of the raw oil
Liquid resin
Lactic acid + Glycolic acid
--* Lactic acid-glycolic acid copolymer
Sorbitol ~ Monoanhydrosorbitol + Dianhydrosorbitol
2,4-Disubstituted phenols +
Formaldehyde ~ Benzodioxin
Formaldehyde ~ Trioxane
M .
. /Benzamide
ethanol +
Substituted hydrazide
[N-Methoxybenzamide
~ Corresponding ester

Lewatit SP 120
(NH~- form)

323

Amberlite IR120B

324

Amberlyst-15

325

idem.

326

idem.

237

Amberlyst-15
Ostion KS
(NH~- form)

328

Amberlite IR-120C
Acidic ion
exchange resin
Strongly acidic
ion exchange resin

330

332

idem.

333

Amberlite IR-120

334

Diaion HPK 55H

335

Dowex HCR-W2-G

336

Ion exchange resin


Amberlite IR-120,
Nation 501
Cation exchange resin

337
338

Amberlyst- 15

340

Transetherification
Miscellaneous

Tetramethoxy silane +
Water + Ethanol --, Silica coating
Aniline + Cyclohexanone
~ N-Cyclohexylideneaniline

Strongly acidic
ion exchange resin
Acidic ion exchange
resin

329

331

339

341
342

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45

21

TABLE 3 (continued)
Reaction

System

Catalyst

Ref.

Rexyn 101H

343

Amberlyst XN-1010

344

Amberlyst-15

345

Lewatit Resin

346

Miscellaneous
1-Hydroxy-4-acyloxy-butane
+ Water ~ Tetrahydrofuran
Acetaldehyde di-ethyl acetal
+ Formamide ~ N-(2-Ethoxyethyl)formamide
1-Butene + Hydrogen sulphide
2-Butanethiol + Di-butyl sulfide
Bisphenol A + 2,6-Dimethylphenol ~ 3,5-Dimethyl4,4'-dihydroxyphenyl-2,2propane + 3,3',5,5'-Tetramethyl-4,4'-dihydroxydiphenyl-2,2-propane +
1,1,3,4,6-Pentamethyl-3-(3,5-dimethyl4-hydroxyphenyl)indan-5-ol
(Phenol
Formaldehyde + { Bisphenol A
~p-tert-Butyl phenol
Phenolic resins
Methyl carbamate + tert-Butanol

N-(tert-Butyl)-methyl-carbamate
Cumene + Oleic acid
Cumylstearic acid
Oleic acid + Acetic acid
Acetoxylation of oleic acid
(Bis(hydroxymethyl)~diarylmethanes
Benzene + |Bis(hydroxymethyB~,diarylethanes
B-Hydroxy ester ~ Fused
and spiro ~,-Butyrolactone
Phthalic anhydride + Phenol
Phenolphthalein
tert-Butyl hydroperoxide
Hydrogen peroxide + Isobutylene
Benzene~ + Benzyl alcohol
Toluene )
[Diphenyl methane
~Benzyl toluene
Substituted acetophenone
1,3,5-Triarylbenzene
2,2'-Dihydroxybiphenyl ~ Dibenzofuran
1A + Acetic anhydride
3,4-Dimethyl-pent-3-ene-2-one
+ 3,4-Dimethyl-pent-4-ene-2-one
DIB + Acetic anhydride
4,6,6-Trimethylhept-3-ene-2-one + 6,6-Dimethylhept-4-ene-2-one

Cation exchange
resin
Perfluorinated
sulfonic acid polymer

347
348

Nation NR-50
Cation exchange resin
Lewatit SP 120

349
350
351

Nation-H

352

Amberlyst-15

353

D61

354

Dowex 50W-XB

355

Nation-H

356

idem.
idem.

357
358

Amberlyst-15
KU-23,
Amberlyst-15

359
360-362

22

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45

TABLE 3 (continued)
Reaction

System

Catalyst

Ref.

Amberlyst-15

362
363

Miscellaneous

Diisoamylene+ Acetic anhydride


--*3,4,5,6,6-Pentamethylhept-3-ene-2-one+ 3,5,6,6Tetramethyl-hept-4-ene-2-one+
3,4,5,6,6,-Pentamethyl-hept-4-ene-2-one
Citronellal ~ Isopulegol
Hydrogenation of cotton seed oil
Cumene + Oxygen ~ Cumene
hydroperoxide
Decomposition of hydrogen
peroxide
2-Methylnaphthalene--*
2-Methyl-l,4-naphthaquinone
Aziridine + Ammonia
Ethylenediamine

many more reactions reported to be catalyzed by ion exchange resins,

10. Reactors used for ion exchange resin


catalyzed reactions
Both slurry reactors and fixed bed reactors
can be used for ion exchange resin catalyzed
reactions. For continuous processes, fixed
bed reactors should be preferred. In fixed
bed reactors, however, small catalyst particles will cause higher pressure drop. Special
type of support trays may also be required to
support very small catalyst particles in fixed
bed reactors,
The use of multitubular reactors packed
with catalyst in the form of particles has
been reported for the manufacture of MTBE
[71]; here the heat is removed by circulation
of water in the shell,

idem.

Ion exchange resin


supported palladium
complexes
Weakly acidic ion
exchange resin
supported cobalt(II)
complexes
Cation exchange
resin supported
copper(II) complexes
Pd-exchangedpolystyrene sulfonic acid
Amberlyst-15

364

365
366
367
368

For MTBE manufacturing, the use of a


distillation column reactor, packed with
cation exchange resin, has been reported [72].
Chem. Res. and Licensing Co., U S A has
suggested a modification in which the upper
portion of the feed is the reactor and the
lower portion is the distillation zone which
ensures a counter-current contact [72]. The
Ethermax process, adopted by U O P and
KOCH for M T B E / T A M E , uses a proprietary structured tower packing developed by
KOCH which acts as an efficient column
packing as well as a catalyst [73].
Enivicerche has claimed that a plate colu m n reactor containing catalyst beds is useful to produce MTBE [74].
CDTECH, U S A has developed a combined etherification/distillation technology
to produce TAME. Catalytic distillation is
effected in a fractionation tower containing
palladium-loaded ion exchange resin catalyst

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45

packing that allows simultaneous catalytic reaction and distillation. Isoamylene conversion of more than 95% has been claimed
[75].
Bubble column reactors have also been
suggested by British Petroleum, UK for
MTBE manufacture [76].
IFP has suggested two process versions for
MTBE manufacturing [77]. The first version
includes an expanded bed front reactor followed by a fixed bed finishing reactor. In the
second version, the expanded bed front reactot is followed by a catalytic distillation reactor as the finishing section. In the first case,
isobutylene conversion is 96-97%, and in the
latter case 98% or more. The expanded bed
reactor is used with some advantages over a
fixed bed reactor or tubular reactor. This
configuration offers inexpensive construction
and provides long catalyst life since no local
overheating takes place, due to the efficient
transfer rate created by the bed expansion
between the catalyst grains and the liquid.
Furthermore it leads to easy loading/
unloading and offers high flexibility with respect to isobutylene concentration and feed
throughput. It is desirable in many cases to
treat the C 4 fraction containing isobutylene
so well that the residual isobutylene is less
than 0.1%. The raffinate C 4 fraction can
then be fractionated to give much needed
and high value added pure 1-butene. (Note,
isobutylene and 1-butene have the same boiling points and cannot be fractionated directly.) Thus a reactor configuration which
meets this requirement is necessary,
A reactive distillation reactor has been
used for alkylation of aromatics [78]. The
cation exchange resin is slurried in the aromatic feed stream and fed to a reaction zone
containing inert distillation packing. The
olefin is vaporized and fed to the bottom of
the reaction zone which agitates the catalyst,
Thus the olefin comes in contact with the
catalyst and the aromatics, leading to the
formation of the alkylated product which is

23

removed from the lower end of the reaction


zone. Any unreacted aromatic is distilled
overhead.
The use of packed bed reactor (packed
with a gel-type cation exchange resin) has
been reported recently by Sakatani and
Shimizu [79] for manufacturing bisphenol A.
The use of loop reactor has been proposed
for reactions using powdered ion exchange
resin as catalyst [80], and its application has
been reported for the oligomerization of
isobutylene [81].

11. Some important observations on ion exchange resin catalysis


The macroreticular resins have proved to
be useful in a variety of acid-catalyzed reactions whereas gel resins are useful for reactions where water is one of the reactants or
products.
A significant and crucial observation in
ion exchange resin catalysis is the non-linear
dependence of reaction rate on the sulfonic
group concentration [4,5,34,64,82,83]. Ancillotti et al. [5] have given a plausible explanation for the above observation. Inside the
resin the hydrogen ion concentration can
reach very high values at which homogeneous catalysis rates are not proportional to
the acid molar concentration but follow
Hammett-type acidity functions. These express the "effective" acid strength of the
medium and increase with an exponent
greater than unity with respect to the hydrogen ion concentration. The values of the
Hammett acidity function, H 0, for some important acid catalysts are given in Table 4.
Other researchers have interpreted their observation in terms of bridging of polar reactants or corresponding carbonium ion into
the hydrogen bonded network of sulfonic
acid groups [34].
The existence of two types of sites in geltype ion exchange resin has been highlighted

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45

24

TABLE 4
Values of Hammett acidity function ( H 0)
Acid

H0

Ref.

Sulphuric acid (100%)

-12.3

369

Fluorosulphonicacid

-15.07

369

p-Toluenesulfonic acid

+ 0.55

(40%)

Amberlyst-15
Nafion
Montmorillonite
Natural

Cation exchanged
HY Zeolites
L a n t h n u m - a n d cerium-

- 2.4

-2.2
-11 to-13
1.5 to - 3 . 0

-5.6 to -8.0
- 1 3 . 6 t o - 12.7
< - 14.5

369
49

370
371
372

372
369
369

exchangedHY zeolites

by Reinicker and Gates [33] in bisphenol-A


synthesis from phenol and acetone in nheptane solvent. The experimental data fitted a model based on the assumption of
competitive adsorption of phenol and acetone on a particular type of site and that of
phenol and n-heptane on another type of
site.
The difference in activity between internal
and external sites of an ion exchange resin
has been reported by different researchers
[84-86]. They have shown with experimental
data and mathematical models that the surface sulfonic acid groups are less active than
those inside the gel phase,
The effectiveness factor data for ion exchange resin catalysts have been reported in
the case of hydration of isobutylene in the
presence of Amberlyst-15 as catalyst by
Leung et al. [87]. They have reported the
effectiveness factor varies from 0.26, for a
1.04 mm diameter particle at 333 K, to 0.84,
for 0.45 mm diameter particle at 303 K. The
microgel effectiveness factor of Amberlyst-15
has been reported to be unity in the hydration of isobutylene by Leung et al. [87] and
Ihm et al. [84].
For liquid/liquid/solid catalyst systems,

such as the hydration of olefins, ion exchange resin catalyst have generally proved
to be ineffective, except in the case of hydration of propylene and butenes. However,
when some co-solvent was used to make the
liquid phase homogeneous, the reaction proceeded smoothly. Panneman and Beenackers
[88-90] have reported different aspects of
cyclohexene hydration with sulfolane as the
co-solvent in the presence of Amberlite XE307. The rate constant depends on the composition of the mixture. The rate constant
decreases between 0 and 60 mol% sulfolane
which is explained by the stabilization of
cyclohexene. Above 60 mol%, the rate constant increases with sulfolane content, which
seems to be due to increased proton activity
with increasing sulfolane content. The equilibrium conversion decreases by a factor of 3
and 6 for solvent mixtures with 60 and 90
mol% sulfolane, respectively. This effect can
be explained quantitatively from the change
in activity coefficients of the reactants and
product with sulfolane content of the mixture. The hydration of a-methylstyrene was
attempted in a liquid/liquid/solid mode
with Amberlyst-15 as the catalyst. No reaction occurred even after several hours of
stirring [91]. However, when isopropanol was
used as co-solvent to make the liquid phase
homogeneous, the reaction proceeded well.
In the etherification of a-methylstyrene
with methanol, using Amberlyst-15 as the
catalyst, the concentration of methanol had a
remarkable effect on the reaction [91]. At
lower concentration of methanol, the rate
showed a negative dependency on the concentration of methanol, whereas at higher
concentration of methanol, the rate was independent of the concentration of methanol.
Macroporous ion exchange resins can be
moulded in the form of Raschig rings, Berl
saddles, etc., and thus can be useful in combined catalysis and separation of the mixture
[92]. The advantages of ion exchange resin
reaction/distillation has been recently re-

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45

ported for the esterification of acetic acid


with ethanol in the presence of Amberlite
IR-120 as a model system [93].
Catalyst-induced yield enhancement for a
reversible reaction has been reported by
Ladisch [94]. The reaction considered was
M T B E manufacture from methanol and
isobutylene and the catalyst used was Amberlyst-35. This appears to contain 1.2-1.8
sulfonic acid group per aromatic ring compared to approximately one sulfonic acid
group per aromatic ring in Amberlyst-15.
Thus the distribution of alcohol in the resin
phase is higher for Amberlyst-35, leading to
a higher mole ratio of alcohol to olefin which
in turn increases the yield.
The ion exchange resin catalyzed process
for manufacturing isopropylchloroacetate offers some prominent advantages over the
homogeneous counterpart, particularly with
reference to the recovery of the unconverted
chloroacetic acid [95]. The method includes
the extraction of the ester in hexane followed
by recrystallization of the unconverted
chloroacetic acid at lower temperature.
The introduction of a polyamide membrane to remove water via pervaporation has
added a new dimension in the manufacturing
of bisphenol A from acetone and phenol in
the presence of cation exchange resin [20%
neutralized with H2N(CH2CH2SH)] as the
catalyst [96,97]. The water formed inhibits
the reaction. The continuous removal of water drives the reaction in the forward direction.
Cation exchange resins can be employed
to separate selectively alkenes or alkynes
from a mixture of gases [98]. The process
comprises contacting of the gas mixture containing alkenes or alkynes with two or more
sequentially arranged adsorbers packed with
H-form of a macroporous strong acidic cation
resin to selectively adsorb one of the alkenes
or alkynes. Water, oxygen or an acid in the
vapour form are then introduced to simultaneously desorb the alkene or alkyne from the

25

resins and to form an alcohol, aldehyde and


ester.

12. Separation of close-boiling compounds


through reactions catalyzed by ion exchange
resin

Separation of isobutylene / 1-butene


The separation of isobutylene from its
mixture with 1-butene (isobutylene and 1butene have almost the same boiling points)
is generally achieved by reacting the mixture
with methanol in the presence of cation exchange resin whereby only isobutylene reacts
[99]. The separation can also be achieved by
selective esterification of isobutylene with
acetic acid in the presence of Amberlyst-15
as the catalyst at 10C [95]. At higher temperature only 1-butene reacts to give secbutyl acetate.

Separationof m-cresol~p-cresol
Separation of m-cresol and p-cresol was
reported by Chaudhuri et al. [100] by reacting a mixture of m-cresol and p-cresol with
AMS or DIB in the presence of Amberlyst-15
as catalyst. At 60C, the reactivity ratio with
AMS as the alkylating agent was 4.8:1 and
that with DIB was 5.6 : 1. The reactivity ratio
was higher in the case of Amberlyst-15 cornpared to that with p-TSA as the catalyst.
Therefore, Amberlyst-15 can be advantageously used for such separations. The corresponding alkylated products were separated
by distillation and cracked to recover the
pure compounds.

Separationof 2,6-xylenol/p-cresol
2,6-Xylenol and p-cresol were separated
from their mixture using the same strategy as
mentioned above [100]. At 60C, in the presence of Amberlyst-15 as the catalyst, the
reactivity ratio with AMS as the alkylating
agent was 13 : 1 and that with DIB was 2.3 : 1.

26

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45

Separation ofphenol/o-chlorophenol, o-ethylphenol/guaiacol


Ghosh and Sharma [7a] have reported the
separation of phenol/o-chlorophenol and
o-ethylphenol/guaiacol by etherification of
the corresponding mixture with isobutylene
or isoamylene in the presence of Amberlyst15 as the catalyst. From a 50 : 50 mixture of
phenol and o-chlorophenol, a 85 : 15 mixture
of tert-butyl ethers of phenol and o-chlorophenol was obtained. At 10C, in presence of
Amberlyst-15 as the catalyst, the reactivity
ratio of o-ethylphenol to guaiacol was found
to be 4 : 1. The corresponding ethers are then
separated by distillation and cracked to recover the pure phenolic compounds separately.

Separation of a-naphthol//3-naphthol
Both a-naphthol and /3-naphthol were
separated in the pure form from a 50:50
mixture via selective etherification with
isobutylene and subsequent selective deetherification [7]. The reactivity ratio of /3naphthol to a-naphthol was reported to be
4:1 in the presence of Amberlyst-15 as the
catalyst.

Separation of 2-nitrobenzaldehyde /3-nitrobenzaldehyde


Separation of 2-nitrobenzaldehyde and 3nitrobenzaldehyde from a 22 : 77 mixture has
been reported by Sainz et al. [101]. The
mixture was acetalized with ethylene glycol
in the presence of IR-120 acidic ion exchange resin as the catalyst. The acetalized
products were separated by crystallization
and subsequent distillation. The separated
acetalized products were then hydrolyzed in
the presence of same catalyst to recover the
pure nitrobenzaldehydes,

has been reported by Hsu [102]. Separation


of 2-methylbutanal from its mixture with
methyl isopropyl ketone was achieved by
treating the mixture with a solid acid catalyst
(Nation). The aldehyde cyclotrimerized in the
presence of Nation, keeping the ketone intact. The ketone and unreacted aldehyde
were distilled off from the mixture. The pure
cyclotrimer was then cracked in the presence
of Nation to recover the aldehyde.

Separationofethanol/isopropanol
Jayadeokar and Sharma [103] have shown
that close-boiling ethanol and isopropanol
can be separated from their mixture by exploiting their reactivity towards isobutylene.
At 50C, in the presence of Amberlyst-15 as
the catalyst, 99% conversion of ethanol to
ethyl tert-butyl ether was observed while at
the same time, conversion of isopropanol to
isopropyl tert-butyl ether was 48.8%. The
tert-butylethers of ethanol and isopropanol
have more than 15C difference in their normal boiling points and hence can be separated.

Separation of isopropanol (IPA) /tert-butanol


(TBA)
IPA and TBA, which are close-boiling, can
be separated using the same strategy as mentioned in the above paragraph. Jayadeokar
and Sharma [103] have reported that at 40C,
in the presence of Amberlyst-15 as the catalyst, IPA reacts with IB while TBA does not
react at all. The difference in normal boiling
points of the two ethers is 19C and hence
can be separated by distillation.

13. Removal of impurities from a mixture by


reaction catalyzed by ion exchange resin

Separation of 2-methylbutanal / methyl isopropyl ketone

Removal of formaldehyde impurity from aqueous solution of butynediol

Separation of aldehyde from a mixture


containing close-boiling aldehyde and ketone

Formaldehyde, a contaminant in the manufacture of 2-butyne-l,4-diol from acetylene

A. Chakrabarti, M.M. Sharma/ React. Polym. 20 (1993) 1-45


and formaldehyde by the Reppe process, is
removed from the aqueous solution of 2butyne-l,4-diol by acetalization with methanol in the presence of Lewatit SPC 118 as
the catalyst [104]. The product methylal is
then removed from the mixture by distillation.

Removal of p-cresol impurity from 2,6-xylenol


The removal of a small amount of p-cresol
impurity from a large excess of 2,6-xylenol
can be achieved by selective etherification of
p-cresol with isobutylene in the presence of
Amberlyst-15 as the catalyst [100]. Thus, a
98:2 (w/w) mixture 2,6-xylenol and p-cresol
could be refined to 99.8 : 0.2 (w/w) by selective etherification of p-cresol with isobutylene at 8C and subsequent distillation of the
resultant mixture,

Removal of a-naphthol impurity from ~naphthol

A small amount of a-naphthol impurity


can be removed from a mixture of a-naphthol and/3-naphthol by exploiting their reactivity difference at 60C towards C-alkylation
[7]. Thus a 95:5 (w/w) mixture of /3-naphthol and a-naphthol could be refined to more
than 99%/3-naphthol by passing isobutylene
through the mixture at 60C in the presence
of Amberlyst-15 as the catalyst and subsequent distillation of the resulting mixture
containing /3-naphthol, a very small amount
of a-naphthol and tert-butyl-o~-naphthol,

14. Modified ion exchange resins and supported perfluorinated resin sulfonic acid

The well-known ion exchange resins like


Amberlyst-15, Lewatit SPC 118, Amberlite
IR-120, etc., cannot be used for extended
periods at temperatures higher than 125C
because of irreversible desulfonation. This
limits the use of conventional ion exchange

27

resins in high temperature processes. This


prompted researchers to think of some modification of the conventional resins to increase
thermal stability. Perfluorosulfonic acid (or
fluorocarbon sulfonic acid as called by some
researchers) which are stable up to 200C
can be used, but cost limits their use in
industrial processes. However, perfluorosulfonic acid can be immobilized on some inert
supports thereby reducing the cost drastically. Recent developments in modified ion
exchange resins and supported perfluorosulfonic acid will be highlighted in the following
section.
The sulfonation procedure described in
section 3 generally gives uniform sulfonation.
However, by deliberately imparting non-uniform sulfonation, remarkable improvement
in selectivity can be achieved for some industrially important reactions [105]. Sulfonations
are defined to be non-uniform if (i) substitution of all the pendant phenyl group is not
achieved or (ii) the pendant phenyl groups
are not primarily para-substituted. Mild sulfonation conditions are applied so that only
the readily accessible macro-phase of the
porous resin is functionalized. By the use of
these resins the selectivity of cracking of
methyl tert-butyl ether (to produce high purity isobutylene for subsequent polymerization) has been significantly improved [105].
Conventional commercial macroporous resins
are not so suitable for the above-mentioned
cracking reaction because they promote the
side reactions, namely, dehydration of methyl
alcohol, oligomerization of isobutylene and
hydration of isobutylene to a relatively larger
extent. The resins obtained from non-uniform sulfonation may also be well suited for
the synthesis of bifunctional hydrogenation
catalysts, since only the required amount of
precious metal, introduced by ion exchange,
will be accessible during hydrogenation.
Resins with higher thermal stability can be
obtained by carrying out the non-uniform
sulfonation at higher temperature (e.g.,

28
160C). The higher temperature favours the
isomerization of para-sulfonic acid groups to
the meta-position leading to an increase in
the thermal stability of the resins. As a certain amount of para-sulfonic acid groups will
remain, the desulfonation of these groups is
observed first in stability tests, with the remaining acid groups being considerably more
stable at 200C. Alternatively, oleum can be
used as a sulfonating agent so that -SO 2bridges are formed during sulfonation, thus
increasing the thermal stability considerably,
The activity and thermal stability of ion
exchange resins can be increased to a great
extent by carrying out acylation of the
polystyrene matrix before sulfonation. Porous
or non-porous polystyrene resins are reacted
with phenylacetyl chloride. After sulfonation
the ortho-acyl derivative shows long-term
stability at 200C in contact with water and
the resin has proved to be highly acidic in
catalytic applications [106].
As an alternative for manufacturing resin
catalysts with improved stability, the synthesis of sulphoalkylated styrene-divinylbenzene copolymer seems to be a very attractive
approach [13]. However, the acidity of the
aliphatic acid groups will be lower, thus restricting its use in mainly aqueous reaction
media, where the effect of acidity is levelled
out as a result of dissociation. These sulphoalkylated resins showed high thermal stability
in water at 200C. Notwithstanding these advantages, the laborious and costly synthesis
of these catalysts will prevent their industrial
application. One of the most attractive methods for further improving the activity and
stability of the resin catalysts is the reaction
of macroporous sulphonated polystyrene
resins with elemental fluorine [107]. The
product was found to be much better as far
as thermal stability and catalytic activity were
concerned. Fast treatment with F 2 at room
temperature produced resins with modified
acid groups in the macro-phase [108]. These
catalysts were more active, especially in

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45

non-polar reactions such as the alkylation of


benzene.
The ion exchange resins can be tailor-made
for different industrially important reactions
by partially exchanging the H ions of the
sulfonic acid resins by different species. Partially mercapto-exchanged resin catalysts are
used in the industrial synthesis of bisphenol
A. Partially Hg 2- or Ag-exchanged resins
are reported to accelerate reactions involving
acetylenic or olefinic compounds, respectively [109]. A trifunctional acid resin containing H , Hg 2 and Fe 3 was used as a
catalyst for acetylene hydration [110]. Moreover, metal-exchanged resins are sometimes
used in the most reactive parts of column
reactors to avoid hot spots and catalyst decomposition in exothermic reactions [22].
The other class of ion exchange resins
with high thermal and chemical stability are
perfluorosulphonic acid polymers, namely,
Nation [111]. These catalysts have higher
acidity as compared to the conventional
resins (see Table 4). The primary advantage
of Nation catalysts does not seem to be the
higher acidity or thermal stability since resins
having the same activity and stability (thermal) can be manufactured from more versatile polystyrene resins, e.g., by fluorination.
The main attraction for using the Nation
catalysts lies in their chemical stability to
strong oxidizing reaction media like H 2 0 2
where the conventional resins fail. Another
attractive property of Nation lies in its ability
to maintain a high acidity even when the
resin is hydrated. In the hydration of ethylene oxide [ethylene oxide:water = 10:90
(w/w)] using Nation coated on Davison silica
952, the conversion of ethylene oxide obtained was 94% with 94% selectivity of ethylene glycol whereas in the case of Amberlyst
XN 1010 (a polystyrene sulfonic acid resin),
the corresponding figures were 19 and 67%,
respectively (other conditions remaining the
same) [112]. Nevertheless, their high price
works against their possible large-scale use in

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45

industry. The cost of perfluorosulfonic acid


can be reduced drastically by coating it onto
inert support. The coating also helps in utilizing maximum possible efficiency by making
maximum number of acidic sites available. It
has been observed by Weaver et al. [113] that
the rate of alkylation of phenol with 1-decene
using 120 mesh powder of perfluorosulfonic
acid resin proceeds at a rate seven times
faster than that obtained with 30 mesh powder. The coating can be done in two ways: (i)
coating with a polymer solution and (ii) coating with a polymer latex. Coating a support
from a polymer solution, followed by heat
treatment produces an improved support catalyst [114]. Coating a support with an aqueous solution of the polymer in the sulfonyl
fluoride form and subsequent conversion of
the sulfonyl fluoride functionality to sulfonic
acid form is more convenient and allows
more flexibility in production of the supported catalyst. In aqueous solution, the size
of the polymer particle is about 100 nm leading to the use of only supports with large
pores (low surface area), 1 ~ m or larger, for
producing supported catalyst from aqueous
solution of the polymer. There is an optimum loading of the polymer on the support
above which the efficiency of the catalyst is
decreased due to pore blockage or excessive
polymer thickness. The optimum polymer
loading for an alumina support is 13 wt%
and that for a silicon carbide support is 20
wt%. The supported perfluorosulfonic acid
resin catalyst showed much higher activity
and working life than the conventional Amberlyst-15. In the alkylation of phenol with
nonene in the presence of Amberlyst-15 conversion of oletin dropped below 50% after 37
days, whereas after the same time, conversion of oletin dropped from 98 to 80% and
remained constant up to 100 days in the
presence of supported catalyst,
The use of supported perfluorinated sulfonic acid resin as the catalyst for oligomerization of oletins, hydration of olefins, etheri-

29

fication of alcohols and hydrolysis of esters


has been recently reported by Knifton [115]
and Martin and Waller [116]. The catalyst
used was Nation supported on calcined shot
coke having hydrophobic surfaces and pore
diameter greater than 1000 ,~.

15. Deactivation and regeneration of ion exchange resin catalysts


Ion exchange resin catalyst may be deactivated as a result of fouling, desulfonation,
chemical deactivation, sulfonic acid neutralization, etc. [117]. The contaminants, solid or
liquid, in the feed can cause fouling. The
active centres might be blocked due to polymerization or polycondensation products.
Desulfonation of the catalyst is temperature
related, and above 110-140C the desulfonation rate becomes significant. Over an extended period, reactants, e.g., oletins, react
with sulfonic groups forming stable compounds and deactivate the catalyst. Metal
cations exchange with the sulfonic acid protons and thus reduce the activity of the catalyst.
The C 4 and C 5 streams from FCC for the
production of M T B E / T A M E may contain
impurities like peroxides, cations, strong
bases, some nitrogen and sulphur compounds and diolefins [118]. Iron in the presence of oxygen can form peroxides which
react with the polystyrene-divinylbenzene
copolymer to separate the aromatic rings
from the polymer chain. This leads to loss of
acid sites on the rings giving reduced catalytic activity and the destruction of the resin
by scission of crosslinks. Cations such as
sodium react with the catalyst acid sites to
replace the hydrogen ion and render the site
inactive. Ammonia and amines neutralize the
sulfonic acid groups and again deactivate the
acidic functionality. Nitriles and amides react
in the presence of the acidic catalyst to produce ammonia and amines which also neu-

30

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45

tralize the catalyst acid function. Dialkyl sultides, under certain conditions, react with
the acid sites and neutralize the catalytic
activity. Diolefins in the C 4 and C 5 feed
streams can oligomerize inside the matrix of
the catalyst thereby fouling the pores of the
catalyst and reducing the catalyst life significantly. However, Mobil has claimed that resin
catalysts can be protected from such damages by carrying out the first stage in a reactor containing a medium pore zeolite where
a major part of the oletin is converted [119].
Dixit and Yadav [120a] have studied the deactivation of ion exchange resin catalysts by
self-poisoning in the alkylation reaction of
o-xylene with styrene at 80C in a mechanically agitated contactor; and they report that
deactivation is a consequence of the formation of styrene oligomer/polymer on the active site of the poly(styrene-co-divinylbenzene (DVB)) support. It was found that the
DVB content plays a major role in the deposition and hence affects the porosity and
pore size distribution of the resin catalyst. As
the DVB content of the catalysts, viz. A m berlyst-16 (12%), Amberlyst-15 (20%), and
Amberlyst-18 (25%), was increased, it was
observed that the porosity reduced by about
5, 8 and 12%, respectively. They have also
carried out simulation studies of the deactivation process with the help of network models. A good agreement between the theoretical and experimental activities was found
[120b].
The literature on regeneration of deactivated ion exchange resin is scanty. The regeneration of the catalyst is possible by removal of the metal ions by acid treatment,
reintroduction of the functional groups or by
treatment with solvents. The regeneration of
Amberlyst-15 used in the synthesis of trioxymethylene from formaldehyde has been
reported by Liu et al. [121]. The poisoning
was caused by coating of macromolecules
formed by condensation polymerization of
formaldehyde and interchange of trace

amounts of metal ions (which were initially


present in the formaldehyde solution) with
protons of the resin. The inorganic poisons
were removed by treating the catalyst with
acid-alkali solutions while the organic poisons were oxidized with nitric acid and subsequently dissolved in caustic soda solution.
The regeneration of ion exchange resin used
in the production of bisphenol A has been
reported by Faler [122]. The destabilizing
impurities are removed from the ion exchanger by treatment with bases and then
with strong acids. The regeneration of deactivated Amberlyst-15 used in the synthesis of
phenylxylylethane from o-xylene and styrene
was tried by Patwardhan and Sharma [123]
by refluxing the catalyst in solvents such as
benzene, toluene, methyl isobutyl ketone, etc.
The regeneration was not realized. It may be
due to the permanent change in the structure of the matrix caused by styrene which is
compatible with the polystyrene matrix. Rehfinger and Hoffmann [61] have reported that
the deactivation of the Amberlyst-15 catalyst
used for the dimerization of isobutene was
caused by blocking of the microparticle gel
phase by higher isobutene oligomers. The
deactivated catalyst could be regenerated
through MTBE synthesis experiment. Novrocik et al. [124] used sulfonated styrene-divinylbenzene copolymer (capacity, 1.15
meq/gmol and mol. wt., 5350) for the esterification of oxalic acid with ethanol. The
patent claims that the regeneration process
used by them gives almost the same activity
as the fresh catalyst.

16. Ion exchange resins


other catalysts

versus

clays versus

Acid-catalyzed reactions are industrially


important and the proper choice of the catalyst is vital to make a process cost-effective.
The acid-catalyzed reactions can be carried
out with both homogeneous and heteroge-

31

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45

neous catalysts. Heterogeneous catalysts have


certain prominent advantages over their homogeneous counterpart, such as easy separation of the catalyst, better selectivity, no corrosion problem, and reusability of the catalyst. Heterogeneous catalysts which are generally employed for acid-catalyzed reactions
are zeolites, ion exchange resins, acidactivated clays, etc. It is not feasible to compare the performance of these catalysts because of lack of experimental data on the
same reaction system using different catalysts. However, in our laboratory we have
carried out some reactions with both ion
exchange resin and acid-activated clay as catalysts. From our observations we will try to
shed light on their performance. Some general remarks about other catalysts will be
made later,
It has been observed that generally the
acid-activated clays are less active compared
to the ion exchange resins. Shah and Sharma
[56] have reported lower activity with acidtreated clay (Engelhard, F-24) than that with
the cation exchange resin (Amberlyst-15) for
the dimerization of isoamylene. The use of
acid-treated clays has improved the selectivity for the desired products in some cases
[125]. In the case of alkylation of phenol with
cyclohexene, selectivity of cyclohexyl phenyl
ether obtained with acid-activated clay was
70% while that with Amberlyst-15 was 59%
(all other conditions remaining the same),
Acid-activated clays have proved to be better
catalysts than ion exchange resins for the
cracking of tert-butyl phenyl ethers [7].
A comparison between montmorillonitederived catalysts and ion exchange resin in
MTBE synthesis has been reported by Adams
et al. [126]. A13+-exchanged montmorillonite
had about half the efficiency of Amberlyst-15
at 60C with dioxane as solvent; however,
K10 or Ti3+-exchanged montmorillonite was
twice as efficient as Amberlyst-15 at the same
temperature. Clays are generally more thermally stable than ion exchange resins. With

conventional ion exchange resins, the higher


temperature limit is less than 150C, whereas
the acid-activated clays are reported to be
useful up to 200C which is the higher temperature limit for perfluorinated sulfonic acid
resin. Pillared clays can withstand much
higher temperature. Clays have the advantage that they are inexpensive and also the
used clays can be adopted as landfill whereas
ion exchange resins which contain sulfur contribute to soil pollution when they are disposed as landfill [127].
The use of supported heteropoly acids
(HPA) in acid-catalyzed reactions is now receiving attention. The use of heteropoly acids
in hydration, etherification, decomposition,
etc., has been reported [128-130]. The
amount of HPA that can be supported on
activated carbon is equal to that of proton
concentration in ion exchange resins [131].
HPA on carbon can be prepared to give a
catalyst having high surface area and high
concentration of active sites on the surface,
thus eliminating the diffusional limitation.
The leaching of HPA limits its use in reactions where polar reactants or solvents are
involved. Titanates are well known as catalysts for esterification. They give high selectivity and better quality product [132]. However, supported titanates are not yet commercially available. Another disadvantage
with the titanates is the requirement of high
temperature in reactions catalyzed by them.

17. Conclusions
The cationic ion exchange resins offer distinct advantages as catalyst for acid-catalyzed
reactions, compared to the use of homogeneous catalysts and other heterogeneous catalysts. A variety of reactions can be carried
out with cationic ion exchange resins as the
catalyst.
Ion exchange resins, moulded in the shape
of tower packings like Raschig rings, can act

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45

32

as packing and catalyst simultaneously. This


can allow distillation concurrently with reaction offering clear advantages in terms of the
energy required per tonne of the product,
Ion exchange resins can be modified to get
a bifunctional catalyst which can be used for
multistep reactions. Resins exchanged with
different metal ions can be used as catalyst
for oxidation, hydrogenation, etc.
A key limitation of cationic ion exchange
resins is the upper temperature limit which is
typically below 120C except in the case of
perfiuorinated resins which can go up to
about 200C. However, these perfluorinated
resins are relatively very expensive. A convenient way to overcome this cost disadvantage
is to adopt supported perfluorinated catalyst,
The demand for environmentally friendly
processes, which are also benign on energy
consumption, and the versatility of cationic
ion exchange resins give them direct advantages in any competition with homogeneous
catalysts and other heterogeneous catalysts,

18. Future scope


Further efforts are required to prepare
supported catalysts and resins which can operate continuously at 160C or above. The
adoption of ion exchange resins as column
packings for distillation reactor systems will
be given further attention.
Efforts are being made to develop perfluorosulfonic acid resin with high surface area
fibres, employing organopolysiloxane substrates rather than polystyrene, to boost their
activity and thermal stability. Modified ion
exchange resin may have the requisite potential for different industrially important reactions. It would be desirable to develop catalysts having large pore sizes so that bulkier
molecules, such as terpenes, can be reacted
via esterification, etherification, hydration,
etc.

The modelling of fixed bed reactor, with

deactivation, merits attention (e.g., for isopropanol from propylene; sec-butanol from
butenes). The role of supercritical conditions
in hydration of propylene and n-butenes
should be systematically studied.
Multifunctional catalysts based on cationic
ion exchange resins should receive more attention.
It would be desirable to carefully select
some more systems of potential industrial
importance where much higher rates are realized with the cationic ion exchange resin
compared to homogeneous catalysts at equivalent acidity level.
The use of cationic ion exchange resin for
removal of impurities in dilute aqueous and
non-aqueous solution merits further attention as this may well provide an environmentally friendly process converting the impurities into useful products.
The regeneration of used ion exchange
resin merits attention. Effective methods of
disposing waste catalysts need to be developed.
The design aspects of expanded-bed reactor and distillation column reactor combinations require investigation. There is considerable scope to exploit ion exchange membrane-reactor configuration where the product of a reaction, such as water, selectively
permeates through the membrane which is
also a catalyst.

References
1 B. Tacke and H. Sfichting, Landwirtsch. Jahrb., 41
(1911) 717~ cf. W. Neir, Ion exchangers as catalysts, in K. Dorfner (Ed.), Ion Exchangers, Walter

de Gruyter, Berlin, 1991.


2 E. Meier and H. Lanth, Organic catalytic reactions, Ger. Pat. 882,091 (1953); Chem. Abstr., 52
(1958) 11121.
3 N.B. Lorette, W.L. Howard and J.H. Brown Jr.,
Preparation of ketone acetals from linear ketones
and alcohols, J. Org. Chem., 24 (1959) 1731.
4 B.C. Gates, J.S. Wisnouskas and H.W. Heath Jr.,

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45


The dehydration of tert-butyl alcohol catalyzed by
sulfonic acid resin, J. Catal., 24 (1972) 320.
5 F. Ancillotti, M.M. Mauri and E. Pescarollo, Ion
exchange resin catalyzed addition of alcohols to
olefins, J. Catal., 46 (1977) 49.
6 S.S. Jayadeokar and M.M. Sharma, Ion exchange
resin catalysed etherification of ethylene and
propylene glycols with isobutylene, React. Polym.,
20 (1993) 57.
7 a. S.K. Ghosh and M.M. Sharma, Separation of
close-boiling phenols and naphthols through their
selective etherification with isobutylene/isoamylene: Acidic ion exchange resin as catalyst, React.
Polym., 16 (1991/1992) 159.
7 b. Anon., Ion Exchange: A new sphere of action,
Chem. Eng., 99(9) (1992) 63.
8 S.S. Bhagade and G.D. Nageshwar, Catalysis by
ion exchange resins - Hydration, Chem. PetroChem. J., 12 (1981) 21.
9 S.S. Bhagade and G.D. Nageshwar, Catalysis by
ion exchange resins - Dehydration, Chem. PetroChem. J., 11 (1980) 23.
10 S.S. Bhagade and G.D. Nageshwar, Catalysis by
ion exchange resins - Esterification, Chem. PetroChem. J., 9 (1978) 3.
11 S.S. Bhagade and G.D. Nageshwar, Catalysis by
ion exchange resins - Ester hydrolysis, Chem.
Petro-Chem. J., 8 (1977) 9.
12 S.S. Bhagade and G.D. Nageshwar, Catalysis by
ion exchange resins - Alkylation, Chem. PetroChem. J., 9 (1978) 21.
13 H. Widdecke, Design and industrial application of
polymeric acid catalysts, in D.C. Sherrington and
P. Hodge (Eds.), Syntheses and Separations using
Functional Polymers, Wiley, Chichester, 1988, p.
149.
14 W. Neier, Ion exchangers as catalysts, in K.
Dorfner (Ed.), Ion Exchangers, Walter de Gruyter,
Berlin 1991, p. 981.
15 G.A. Olah, P.S. Iyer and G.K.S. Prakash, Perfluorinated resin sulfonic acid (Nation-H) catalysis in
synthesis, Synthesis, (1986) 513.
16 F. Heifferich, Ion Exchange, McGraw-Hill, New
York, 1962.
17 I.M. Abrams, High porosity polystyrene cation exchange resins, Ind. Eng. Chem., 48 (1956) 1469.
18 K.W. Pepper, Sulphonated crosslinked polystyrene: A monofunctional cation-exchange resin, J.
Appl. Chem., 1 (1951) 124.
19 R.M. Wheaton and D.F. Harrington, Preparation
of cation exchange resins of high physical stability,
Ind. Eng. Chem., 44(1952)1796.
20 W.P. Hohenstein and H. Mark, Polymerization of

33
olefins and diolefins in suspension and emulsion.
Part I, J. Polym. Sci., 1 (1946) 127.
21 F.H. Winslow and W. Matreyek, Particle size in
suspension polymerization, Ind. Eng. Chem., 43
(1951) 1108.
22 Permutit Co. Ltd. and J.R. Millar, Brit. Pat.
849,122 (1960); cf. F. Helfferich, Ion Exchange,
McGraw-Hill, New York, 1962, p. 60.
23 P.M. Lange and F.B. Martinola, Monodisperse ion
exchangers and adsorbents: New variations in exchange technology, in M. Streat (Ed.), Ion Exchange for Industry, Ellis Horwood, Chichester,
1988.
24 J.D. McClure, Ethylbenzene process using an unsupported perfluorinated polymer catalyst, US
4,041,090 (1977); Chem. Abstr., 87 (1977) 184185.
25 A.R. Pitochelli, Ion Exchange Catalysis and Matrix
Effects, Rohm and Haas, Philadelphia, 1980.
26 H. Corte and A. Meyer, Anion exchangers with a
spongy structure, Ger. Pat. 1,045,102 (1958); Chem.
Abstr., 55 (1961) 1969.
27 H. Corte, E. Meier and H. Seifert, Cation exchange resins prepared by non-branching polymerization of cross-linked polymers, Ger. Pat.
1,113,570 (1957); Chem. Abstr., 56 (1962) 6173.
28 R. Kunin, E. Meitzner and N. Bortnick, Macroreticular ion exchange resins, J. Am. Chem. Soc., 84
(1962) 305.
29 R.E. Kressman and J.R. Millar, Ion-exchange processes, Brit. Pat. 889,304, (1962); Chem. Abstr., 57
(1962) 1096.
30 Rohm and Haas Co., Cross-linked copolymers for
use in ion exchange resins, Brit. Pat. 932,125 and
932,126 (1963); Chem. Abstr., 59 (1963) 11731.
31 N.M. Bortnick, Catalyzing reactions with cationexchange resins, US Pat. 3,037,052 (1962); Chem.
Abstr., 57 (1962)6125.
32 C.J. Litteral, Rearrangement of organosiloxanes
using macroreticular sulfonic acid cation-exchange
resin, US Pat. 3,694,405 (1972); Chem. Abstr., 78
(1973) 17047.
33 R.A. Reinicker and B.C. Gates, Bisphenol A synthesis: Kinetics of the phenol-acetone condensation reaction catalyzed by sulfonic acid resin,
AIChE J., 20 (1974) 933.
34 R.B. Wesley and B.C. Gates, Benzene propylation
catalyzed by sulfonic acid resin, J. Catal., 34 (1974)
288.
35 G. Naumann, Catalysis by ion exchangers, Chem.
Tech., 11 (1959) 18.
36 F. Helfferich, Katalyse durch Ionenaustauscher,
Ein Hilfsmittel der pr~iparativen organischen
Chemic. Angew. Chem., 66 (1954) 241.

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45

34
37 C.W. Davies and G.G. Thomas, Ion exchange
resins as catalysts in the hydrolysis of esters, J.
Chem. Soc., (1952) 1607.
38 V.C. Haskell and L.P. Hammett, Rates and temperature coefficients in the hydrolysis of some
aliphatic esters with a cation exchange resin as the
catalyst, J. Am. Chem. Soc., 71 (1949) 1284.
39 G. Zundel, Hydration and lntermolecular Interac-

40

41

42

43

44
45

46

47

48

49

50

51

52

tion Infrared Investigations with Polyelectrolyte


Membranes, Academic Press, New York, 1969.
R. Keller, H. Zinke-Allmang, E. Keyssner and P.
Pfaff, Columns for performing reactions in the
presence of catalyst beads, DE 1075613 (1957);
Chem. Abstr., 55 (1961)15018.
W. Neier, Direct hydration of propylene, in D.
Naden and M. Streat (Eds.), Ion Exchange Technology, Ellis Horwood, Chichester, 1984, p. 560.
H. Hasegawa and T. Higashimura, Selective alkyiation of aromatic hydrocarbons with styrene by
solid polymeric oxo acids, Polym. J., 12 (1980) 407.
B. Chaudhuri and M.M. Sharma, Alkylation of
phenol with a-methylstyrene, propylene, butenes,
isoamylene, 1-octene and diisobutylene, Ind. Eng.
Chem. Res., 30 (1991) 227.
K. Gokul Chandra and M.M. Sharma, unpublished work.
A. Guyot, P. Hodge, D.C. Sherrington and H.
Widdecke, Recent studies aimed at the development of polymer-supported reactants with improved accessibility and capacity, React. Polym.,
16 (1991/1992) 233.
L.K. Doraiswamy and M.M. Sharma, Heterogeneous Reactions: Analysis, Examples, and Reactor
Design, Vol. 2, Wiley-Interscience, New York,
1984, p. 254.
L.K. Doraiswamy and M.M. Sharma, Heterogeneous Reactions: Analysis, Examples, and Reactor
Design, Vol. 2, Wiley-Interscience, New York,
1984, P. 280, p. 307.
S. Carra, Reaction processes involving ion-exchange resins, in A.E. Rodrigues (Ed.), Ion Exchange: Science & Technology, Martinus Nijhoff,
Dordrecht, 1986.
C. Buttersack, Accessibility and catalytic activity
of sulfonic acid ion exchange resins in different
solvents, React. Polym., 10 (1989) 143.
S.-K. Ihm, S.-S. Suh and I.-H. Oh, Reaction and
mass transfer in a macroreticular resin catalyst, J.
Chem. Eng. Jpn., 15 (1982) 206.
T. Yoshioka and M. Shimamura, Studies of
polystyrene-based ion-exchange fiber. II. Catalyst
for sucrose inversion and methyl acetate hydrolysis, Bull. Chem. Soc. Jpn., 57 (1984) 334.
L. Petrus, R.W. De Roo, E.J. Stamhuis and G.E.H.

53

54
55

56

57

58

59

60

61

62

63

64

65

66

Joosten, Kinetics and equilibria of the hydration


of propene over a strong acid ion exchange resin
as catalyst, Chem. Eng. Sci., 39 (1984) 433.
L. Petrus, R.W. De Roo, E.J. Stamhuis and G.E.H.
Joosten, Kinetics and equilibria of the hydration
of linear butenes over a strong acid ion exchange
resin as catalyst, Chem. Eng. Sci., 41 (1986) 217.
B.C. Gates, Catalytic Chemistry, Wiley, New York,
1992.
W.O. Haag, Oligomerization of isobutylene on
cation exchange resin, Chem. Eng. Prog. Symp.
Ser., 63(73)(1967)140.
N.F. Shah and M.M. Sharma, Dimerization of
isoamylene: Ion exchange resin and acid-treated
clay as catalysts, React. Polym., 19 (1993) 181.
M. Voloch, M.R. Ladisch and G.T. Tsao, Methyl
t-butyl ether (MTBE) process catalyst parameters,
React. Polym., 4 (1986) 91.
C. Subramaniam and S. Bhatia, Liquid phase synthesis of methyl tert-butyl ether catalyzed by ion
exchange resin, Can. J. Chem. Eng., 65 (1987) 613.
A. Rehfinger and U. Hoffmann, Kinetics of methyl
tertiary butyl ether liquid phase synthesis catalyzed by ion exchange resin. I. Intrinsic rate expression in liquid phase activities, Chem. Eng. Sci.,
45 (1990) 1605.
A. Rehfinger and U. Hoffmann, Kinetics of methyl
tertiary butyl ether liquid phase synthesis catalyzed by ion exchange resin. II. Macropore diffusion of methanol as rate-controlling step, Chem.
Eng. Sci., 45 (1990) 1619.
A. Rehfinger and U. Hoffmann, Formation of
di-isobutene, Main by-product of methyl tertiary
butyl ether synthesis catalyzed by ion exchange
resin, Chem. Eng. Technol., 13 (1990) 150.
K. Sundmacher and U. Hoffmann, Importance of
irreversible thermodynamics for liquid phase ion
exchange catalysis: Experimental verification for
MTBE-synthesis, Chem. Eng. Sci., 47 (1992) 2733.
A. Gicquel and B. Torck, Synthesis of methyl
tertiary butyl ether catalyzed by ion-exchange resin.
Influence of methanol concentration and temperature, J. Catal., 83 (1983) 9.
B.C. Gates and L.N. Johanson, The dehydration
of methanol and ethanol catalyzed by polystyrene
sulfonate resins, J. Catal., 14 (1969) 69.
F. Cunill, L. Toral, J.F. Izquiordo, J. Tejero and
M. Iborra, Influence of resin type and water on
the kinetics of the decomposition of methyl tertbutyl ether in the gas phase, React. Polym., 10
(1989) 175.
M. Iborra, J.F. Izquierdo, F. Cunill and J. Tejero,
Application of the response surface methodology
to the kinetic study of the gas-phase addition of

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45


ethanol to isobutene on a sulfonated styrene-divinylbenzene resin, Ind. Eng. Chem. Res., 31(8)
(1992) 1840.
67 C. Buttersack, H. Widdecke and J. Klein, Sulfonic
acid ion-exchange resins as catalysts in non-polar
media. II. React. Polym., 5 (1987) 181.
68 M.M. Sharma, Perspectives in gas-liquid reactions, Chem. Eng. Sci., 38(1) (1983) 21.
69 F. Helfferich, A quantitative approach to ion exchange catalysis, J. Am. Chem. Soc., 76 (1954)
5567.
70 L. Woliotis, Thesis, G6ttingen, (1958); cf. F. Helfferich, Ion Exchange, McGraw-Hill, New York,
1962.
71 J.C. Davis and M.K. Philip, MTBE Bandwagon,
Chem. Eng., 86(11)(1979)91.
72 L.A. Smith Jr., Process and apparatus for the
preparation of methyl tertiary butyl ether, US Pat.
4,978,807 (1990); Chem. Abstr., 114 (1991) 145832.
73 UOP and KOCH, Upgrade MTBE/TAME Technoiogy, Eur. Chem. News, (1991), June 3, p. 24.
74 D. Sanfilippo, M. Lupieri and F. Ancilloti, Preparation of tertiary alkyl ethers and apparatus for
reactive distillation, EP 470,655 (1992); Chem. Abstr., 116 (1992) 237789.
75 K. Rock, TAME: Technology merits, Hydrocarbon
Process., 71(5) (1992) 86.
76 K. Allott, MTBE takes the knocks, Process Eng.,
72(6) (1991) 50.
77 MTBE, ETBE, TAME; IFP etherification technology, Inst. Fr. Petrole-Direction Ind. Bull., (1991)
12.
78 L.A. Smith Jr. and J.R. Adams, Process for the
alkylation of organic aromatic compounds, Eur.
Pat. Appl. EP 387,080 (1990); Chem. Abstr., 114
(1991) 64757.
79 T. Sakatani and H. Shimizu, Preparation of
bisphenol A, Jpn. Kokai Tokkyo Koho JP 0401,149
(1992); Chem. Abstr., 116 (1992) 255307.
80 F. Martinola, Ion exchangers and adsorbents Versatile aids for the chemical industry', Ger.
Chem. Eng., 3 (1980) 79.
81 W. Kroenig and G. Scharfe, Erdoel Kohle, 19
(1966) 497; cf. D.C. Sherrington and P. Hodge
(Eds.), Synthesis and Separation using Functional
Polymers, Wiley, Chichester, 1988, p. 161.
82 T. Uematsu, Isomerization of butenes over acidic
ion-exchange resins, Bull. Chem. Soc. Jpn., 45
(1972) 3329.
83 B.C. Gates and G.-M. Schwab, The dehydration of
formic acid catalyzed by polystyrene sulfonic acid,
J. Catal., 15 (1969) 430.
84 S. Ihm, M. Chung and K. Park, Activity difference
between the internal and external sulfonic groups
of macroreticular ion-exchange resin catalysts in

35

85

86

87

88

89

90

91

92

93

94

95

96

97

isobutylene hydration, Ind. Eng. Chem. Res., 27


(1988) 41.
S.-K. Ihm and J.-H. Ahn, Interaction of reaction
and diffusion in macroreticular sulfonic acid resin
catalyst, in K. Tanabe, H. Hattori, T. Yamaguchi
and T. Tanaka, (Eds.), Acid-Base Catalysis, Kodansha, Tokyo, 1989, p. 467.
K.M. Dooley, J.A. Williams, B.C. Gates and R.L.
Albright, Sulfonated poly(styrene-divinylbenzene)
catalysts. 1. Diffusion and the influence of macroporous polymer physical properties on the rate of
reesterification, J. Catal., 74 (1982) 361.
P. Leung, C. Zorrilla, F. Recasens and J.M. Smith,
Hydration of Isobutene in liquid-full and tricklebed reactors, AIChE J., 32 (1986) 1839.
H.-J. Panneman and A.A.C.M. Beenackers, Solvent effects on the hydration of cyclohexene catalyzed by a strong acid ion exchange resin. 1.
Solubility of cyclohexene in aqueous sulfolane
mixtures, Ind. Eng. Chem. Res., 31 (192) 1227.
H.-J. Panneman and A.A.C.M. Beenackers, Solvent effects on the hydration of cyclohexene catalyzed by a strong acid ion exchange resin. 2.
Effect of sulfolane on the reaction kinetics, Ind.
Eng. Chem. Res., 31 (1992) 1425.
H.-J. Panneman and A.A.C.M. Beenackers, Solvent effects on the hydration of cyclohexene catalyzed by a strong acid ion exchange resin. 2.
Effect of sulfolane on the equilibrium conversion,
Ind. Eng. Chem. Res., 31 (1992) 1433.
A. Chakrabarti and M.M. Sharma, Ion exchange
resin catalyzed hydration of a-methylstyrene and
etherification of a-methyl-styrene with methanol,
React. Polym., 18 (1992) 117.
K. Gottlieb, W. Graf, K. Schaedlich, U. Hoffmann, A. Rehfinger and F. Joerg, Macroporous
ion exchanger moldings and their uses, Eur. Pat.
Appl. EP 417,407 (1991); Chem. Abstr., 115 (1991)
30781.
J. Savkovic-Stevanovic, M. Misic-Vukovic, G. Boncic-Caricic, B. Trisovic and S. Jezdic, Reaction
distillation with ion exchangers, Sep. Sci. Technol.,
27 (1992)613.
M.R. Ladisch, Catalyst-induced yield enhancement of reversible chemical processes, PCT Int.
Appl. WO 90 08,758, (1990); Chem. Abstr., 114
(1991) 101127.
A.A. Patwardhan and M.M. Sharma, Esterification of carboxylic acids with olefins using cation
exchange resins as catalysts, React. Polym., 13
(1990) 161.
K. Okamoto, T. Semoto, K. Tanaka and H. Klta,
Application of pervaporation to phenol acetone
condensation reactions, Chem. Lett., (1991), 167.
K. Okamoto, H. Kita, Y. Tanaka and S. Iimuro,

36
Method for preparing bisphenol A, EP 442,122
(1991); Chem. Abstr., 115 (1991) 158713.
98 M. Kling, Process for the selective separation of
alkenes and alkynes, US Pat. 4,992,601 (1991);
Chem. Abstr., 114 (1991) 184775.
99 A. Convers, B. Jugin and B. Torck, Make-up
butenes via MTBE, Hydrocarbon Proccess., 60
(1981) 95.
100 B. Chaudhuri, A.A. Patwardhan and M.M.
Sharma, Alkylation of substituted phenols with
olefins and separation of close-boiling phenolic
substances via alkylation/dealkylation, Ind. Eng.
Chem. Res., 29 (1990) 1025.
101 D. Sainz, I. Claro and J. Santa-Olalla, Process for
selective separation of nitrobenzaldehyde isomers
via diol acetals, Span ES 2,020,891 (1991); Chem.
Abstr., 116 (1992) 214144.
102 W.L. Hsu, Separation of aldehydes from ketone
via acid catalyzed cyclotrimerization of the aidehyde, US Pat. 4,701,561 (1987); Chem. Abstr., 108
(1988) 40094.
103 S.S. Jaydeokar and M.M. Sharma, Absorption of
isobutylene in aqueous ethanol and mixed alcohols: Cation exchange resins as catalyst, Chem.
Eng. Sci., 47(13/14) (1992) 3777.
104 F. Merger and H. Horler, Process for removing
formaldehyde from aqueous solutions of 2-butyne1,4-diol, Eur. Pat. Appl. EP 309,915 (1989); Chem.
Abstr., 111 (1989) 194103.
105 U. Haupt, H. Widdecke and J. Klein, Pure tertiary
olefins, Ger. Pat. 3,509,292 (1985); Chem. Abstr.,
104 (1986)148292.
106 K. Struss and H. Widdecke, Strong-acid cation
exchangers, Ger. Pat. 3344210 (1987); Chem. Abstr., 107 (1987) 155423.
107 J. Klein, H. Widdecke, F. Doescher and F. Pohl,
Fluorination of polymers and perfluorinated ion
exchange resin, Ger. Pat. 3,023,455 (1982), Chem.
Abstr., 96 (1982)104970.
108 J. Schlfiter, Ph.D. thesis, Technical University,
Braunschweig, 1987; cf. H. Waddecke, Design and
"industrial application of polymeric acid catalysts,
in D.C. Sherrington and P. Hodge (Eds.), Separations using Functional Polymers, Wiley, Chichester,
1988.
109 F.J. Waller, Catalysis with metal cation-exchange
resins, Catal. Rev. Sci. Eng., 28 (1986) 1.
110 T.T. Moxley and B.C. Gates, A trifunctional ionexchange resin catalyst for acetylene hydration, J.
Mol. Catal., 12 (1981)389.
111 F.J. Waller, Catalysis with metal cation-exchange
resins, ACS Symp. Ser., 42 (308) 1986.
112 L. Kim, Catalytic hydration of ethylene oxide to

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45


ethylene glycol, US Pat. 4,165,440 (1979); Chem.
Abstr., 91 (1979) 158329.
113 J. Weaver, E. Tasset and W.E. Fry, Supported
fluorocarbonsulfonic acid polymer heterogeneous
acid catalyst, Stud. Su E Sci. Catal., 38 (1988) 484.
114 J.D. Weaver, E.L. Tasset and W.E. Fry, Supported Fluorocarbonsulfonic Acid Catalysts, Catal.
Today, 14 (1992) 195.
115 J.F. Knifton, One step synthesis of methyl tertbutyl ether from tert-butanol using fluoro sulphonic acid-modified zeolite catalysts, US Pat.
5,081,318 (1992); Chem. Abstr., 116 (1992) 105622.
116 H.D.B. Martin and F.J. Waller, Supported perfluorinated ion-exchange polymers as catalysts, US
Pat. 5,094,995 (1992); Chem. Abstr., 116 (1992)
195946.
117 H.L. Brockwell, P.R. Sarathy and R. Trotta, Synthesize ethers, Hydrocarbon Process., 70(9) (1991)
133.
118 R. Gonzalez, Some views on clean fuels technology, Hydrocarbon Process., 71(1) (1992) 21.
119 M.N. Handi and H. Owen, Protection of
polystyrene sulfonic acid resin catalyst in a multistage process for preparing unsymmetrical tert-alkyl ethers, US Pat. 5,015,782 (1991): Chem. Abstr.,
115 (1992)91660.
120 a. A.B. Dixit and G.D. Yadav, Deactivation of ion
exchange resin catalyst in alkylation reaction of
o-xylene with styrene. I. Pore structure characterization, to be published.
120 b. A.B. Dixit and G.D. Yadav, Deactivation of ion
exchange resin catalyst in alkylation reaction of
p-xylene with styrene. II. Simulation by network
models, to be published.
121 H. Liu, K. Yang and F. Li, Study of ion exchange
resin catalyst regenerating method in the synthesis
of trioxymethylene, Lizi Jiaohuan Yu Xifu, 4 (1988)
298 (in Chinese); Chem. Abstr., 110 (1989) 233560.
122 G.R. Faler, Method for purification of ion exchange resins used in production of bisphenol A,
EP 324,080 (1989); Chem. Abstr., 111 (1989)
233860.
123 A.A. Patwardhan and M.M. Sharma, unpublished
Work.
124 J. Novrocik, M. Malik, M. Novrocikova, P.
Proksova and M. Zemanova, Catalyst for esterification of oxalic acid, Czech. CS 265,260 (1990);
Chem. Abstr., 117 (1992) 47936.
125 A. Chakrabarti and M.M. Sharma, Alkylation of
phenol with cyclohexene catalyzed by cationic
ion-exchange resins and acid-treated clay: O-versus
C-Alkylation, React. Polym. 17 (1992) 331.
126 J.M. Adams, K. Martin, R.W. McCabe and S.

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45


Murray. Methyl tert-butyl ether (MTBE) production: A comparison of montmorillonite-derived
catalysts with an ion exchange resin, Clays Clay
Miner., 34 (1986) 597.
127 J.M. Thomas, Solid acid catalysts, Sci. Am., 266(4)
(1992) 82.
128 J.F. Knifton and J.R. Sanderson, Heteropoly acid
and sulfonated ion exchange resin catalysts for
production of phenol and acetone from cumene
hydroperoxide, US Pat. 4,898,995 (1990); Chem.
Abstr., 113 (1990) 131728.
129 T. Yamada and T. Muto, Development of new
hydration process of 1-butene, Sekiyu Gakkishi,
34(3) (1991) 201; Chem. Abstr., 115 (1991) 52226.
130 J.F. Knifton, Synthesis of low molecular weight
ethylene and propylene glycol ethers via olefin
addition to the corresponding glycol, EP 419,077
(1991); Chem. Abstr., 115 (1991).48848.
131 M.A. Schwegler, P. Vinke, M. van der Eijk and H.
van Bekkum, Activated carbon as a support for
heteropolyanion catalysts, Appl. Catal. A, 80(1992)
41.
132 C. Blandy, D. Gervals, J.-L. Pellegatta, B. Gilot
and R. Guiraud, Stearic acid esterification catalyzed by titanates: Comparative study of homogeneous and heterogeneous systems, J. Mol. Catal.,
64 (1991) L1.
133 F.W. Melpolder and W.W. Wentzheimer, Tertiary
butyl methyl ether, US Pat. 4198530 (1980); C h e m .
Abstr., 93 (1980) 113957.
134 B. Vora, P. Pujado, T. Imai and T. Fritsch, Production of detergent olefins and linear alkylbenzenes, Chem. Ind., 19 (1990) 187.
135 F. Obenaus, B. Greving, H. Balke and B. Scholz,
High-purity isobutene by dehydration of tertbutanol, DE 3,151,446 (1983); Chem. Abstr., 99
(1983) 123116.
136 G.H. Beasley and I.J. Jakovac, Ion exchange resin
catalysts having improved catalytic activity and enhanced thermal stability, in D. Naden and M.
Streat (Eds.), Ion Exchange Technology, Horwood,
Chichester, 1984, p. 440.
137 Chem. Econ. Eng. Rev., 15 (May 1983), p. 16; cf.
J.N. Armor, New catalytic technology commercialized in the USA during the 1980's, Appl. Catal.,
78 (1991) 141.
138 G. Brandes, W. Neier, J. Woellner, W. Webers
and A. Hoppe, Continuous manufacture of isopropyl alcohol, DE 21 47,739 (1973); Chem. Abstr.,
78 (1973) 158900.
139 G. Brandes, W. Neier, J. Woellner and W. Webers, Continuous manufacture of isopropyl alcohol, DE 21 47,740, (1973); Chem. Abstr., 78 (1973)
158899.
140 G. Brandes, W. Neier, J. Woellner, W. Webers

37
and W.F. De Viesschauwer, Continuous preparation of 2-propanoi, DE 22 33,967 (1974); Chem.
Abstr., 80 (1974) 120238.
141 W. Webers, L. Sandhack and W. Neier, Alcohols
by direct catalytic hydration of olefins, DE 24
29,770 (1976); Chem. Abstr., 84 (1976) 121114.
142 W. Neier, W. Webers, R. Ruckhaber, G. Osterburg and W. Ostwald, Continuous production of
sec-butanol, DE 30 40,997 (1982); Chem. Abstr., 97
(1982) 25525.
143 E.I. Leupold and H.J. Arpe, Isoporopyl carboxylates, Ger. Often. 2,306,586 (1974); Chem. Abstr.,
81 (1974)135472.
144 F. Ancillotti, S.D. Gioacchino and G. Costa, Secondary butanol, Ger. Often, 3,003,126 (1980);
Chem. Abstr., 94 (1981)30174.
145 M.A. Perez Pascual and F. de Benito Albertos,
Process for obtaining sec-butyl acrylate, Eur. Pat.
Appl. EP 445,859 (1991); Chem. Abstr., 115 (1991)
233104.
146 E. Lohma, A. Ohorodink, K. Gehrmann and P.
Stutzke, Esterification of acrylic acid, Ger. Often,
2,226,829 (1973); Chem. Abstr., 80 (1974)70363.
147 H. Erpenbach, K. Gehrmann, H. Joest and P.
Zerres, Dibutyl ether-free n-butyl acrylate Ger.
Often. 2,449,811 (1976); Chem. Abstr., 85 (1976)
47301.
148 H. Matsumura, F. Murakami and H. Sonobe,
Methacrylic acid esters, Eur. Pat. Appt. 10,953
(1980); Chem. Abstr., 93 (1980) 133070.
149 F. Helmut, Phenol derivatives, in B. Elvers, S.
Hawkins and G. Schulz (Eds.), Ullmann's Encyclopedia of Industrial Chemistry, Vol. 19, VCH, Weinheim, 1991, p. 313.
150 Anon., Methyl Glucoside A-Plenty, Chem. Eng.
News, Oct 24, (1955), p. 4592.
151 Rheinpreussen AG Fi~r Bergbau und Chemie,
Methyl isopropenyl ketone, Fr. Pat. 1,383,548
(1964); Chem. Abstr., 62 (1965) 9013.
152 J. Woellner and W. Neier, Continuous one step
preparation of saturated carbonyl compounds,
Ger. Offen. 1,260,454 (1968); Chem. Abstr., 69
(1968) 35439.
153 A. Blanc, F. Hamedi-Sangsari and F. Chastrette,
Process for the preparation of glyoxal monoacetals, Eur. Pat. Appl. EP 249,530 (1987); Chem.
Abstr., 109 (1988)22537.
154 R. Wei and R. Che, Synthesis of furfural glycol
acetal with ion exchange resin as catalyst, Lizi
Jiaohuan Yu Xifu, 7 (1991) 202 (in Chinese); Chem.
Abstr., 116 (1992) 176452.
155 G.A. Olah, S.C. Narang, D. Meider and G.F.
Salem, Catalysis by solid superacids; 8. Improved
Nafion-H perfluorinated resin sulfonic acid-cataiyzed preparation of dimethyl acetals and eth-

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45

38

156

157

158

159

160

161

162

163

164

165

166

167

168

ylenedithioacetals, and hydrolysis of dimethyl acetals, Synthesis, (1981)282.


A.E. Dann, J.B. Davis and M.J. Nagler, A rapid
and convenient technique for converting ketones
into their ethylenedioxy or trimethylenedioxy
derivatives and for making acetonides, J. Chem.
Soc., Perkin Trans. I, (1979) 158.
G.A. Olah, J. Kaspi and J. Bukala, Heterogeneous
catalysis by solid superacids. 3. Alkylation of benzene and transalkylation of alkylbenzenes over
graphite-intercalated Lewis acid halide and perfluorinated resin-sulfonic acid (Nation-H) catalysts,
J. Org. Chem., 42 (1977) 4187.
H.-S. Park and S.-K. Ihm, Alkylation of benzene
with 1-dodecene by macroreticular resin catalysts,
Korean J. Chem. Eng., 2 (1985) 69.
T. Oe, M. Matsuo, K. Yamaguchi and T. Yamaguchi, Preparation of indan derivatives, Jpn.
Kokai Tokkyo Koho JP 03,190,830 (1991); Chem.
Abstr., 116 (1992)83272.
Y. Shirato, S. Asaoka, H. Tajima, Y. Tachibana,
K. Tate and H. Taniguchi, Preparation of a-Isopropyl naphthalenes, Jpn. Kokai Tokkyo Koho JP,
0288,533, (1990); Chem. Abstr., 113 (1990) 58716.
H. Widdecke and J. Klein, Alkylation of phenol
with gaseous olefins on polymeric solid acids,
Chem. Ing. Tech., 53 (1981) 954.
H. Airs, Simultaneous production of o-, p-, and
2,4-di-tert-butylphenol, Ger. Often. DE 3,443,736
(1986); Chem. Abstr., 105 (1986) 114730.
K. Zieborak, W. Ratajczak and H. Kowalska, Synthesis of p-cumylphenol in the presence of a cation
exchanger as a catalyst, Chem. Stosow., 26 (1982)
341.
H. Alfs, W. Boexkes and E. Vangermain, p-tertOctylphenol by catalytic alkylation of phenol, Ger.
Often, DE 3,151,693 (1983); Chem. Abstr., 99
(1983) 139496.
A.A. Patwardhan and M.M. Sharma, Alkylation of
phenol with 1-dodecene and diisobutylene in the
presence of a cation exchanger as the catalyst,
Ind. Eng. Chem. Res., 29 (1990) 29.
P.S. Belov, E.N. Grigor'eva and O.N. Tsvetkov,
Kinetics of phenol alkylation by higher olefins in
the presence of cation exchange resins, Khim.
Tekhnol. Toplic Masel, 7 (1981) 11.
A.G. Trubnikov, V.A. Zavorotngi, S.K. Shakirov,
P.P. Kapustin, A.A. Gritsenko, Y.A. Rumyantsov,
V.K. Khaliullin, P. Belov and K.D. Korenev,
Method of preparation of alkyl phenols, SU
1,703,637 (1992); Chem. Abstr., 117 (1992) 69562.
E. Grzywa, J. Polaczek, M. Kiedik, Z. Lisicki, G.

169

170

171

172

173
174

175

176

177

178

179

180

181

182

Collin and A. Alscher, Process for preparation of


alkyl- or bisphenols, Ger. Often. DE 4,014,992
(1991); Chem. Abstr., 116 (1992) 41081.
Y. Naruse, and S. Sonobe, Process for the catalytic
preparation of monoaralkylphenols from phenols
and styrenes, Jpn. Kokai Tokkyo Koho JP 63
139,142 (1988); Chem. Abstr., 110 (1989) 114441.
S. Sonobe and Y. Naruse, Preparation of monoaralkylphenols, Jpn. Kokai Tokkyo Koho JP
63,227,533 (1988); Chem. Abstr., 110 (1989) 38741.
B. Hoffmann, Z. Witte and F. Wika, Preparation
of 2-methyl-2-(4-hydroxyphenyl)butane, Pol. PL
152,870 (1991); Chem. Abstr., 116 (1992) 105802.
Mitsui Petrochemical Industries Ltd., Alkylphenols, Jpn. Kokai Tokkyo Koho JP 58 83,640 (1983);
Chem. Abstr., 99 (1983) 160306.
R.M. Parlman, Alkyl aryl ethers, US 4,299,996
(1981); Chem. Abstr., 96 (1982) 51995.
E. Santacesaria, R. Silvani, P. Wilkinson and S.
Carra, Alkylation of p-cresol with isobutene catalyzed by cation-exchange resins: A kinetic study,
Ind. Eng. Chem. Res., 27 (1988) 541.
M.C. Crozat, J.M. Barriac, L.T. Aikawa and L.
Bornholdt, tert-Butyl phenols, Braz Pedido PI BR
80 02 607 (1981); Chem. Abstr., 96 (1982) 142455.
W. Hanbold, K. Seiffarth and U. Gladigan, Mono
-tert-butylation of cresols, DD 227,960 (1985);
Chem. Abstr., 105 (1986)6309.
S.D. Kim, H.K. Lee and Y.G. Kim, A study on the
tert-butylation of hydroquinones, Hwahak Konghak, 30(1) (1992) 18; Chem. Abstr., 117 (1992)
89901.
F. Hiroshi, O. Matsumoto and F. Kawaguchi,
Manufacture of tertiary butylated poly(p-vinylphenol), Jpn. Kokai Tokkyo Koho JP 63,154,704
(1988); Chem. Abstr., 109 (1988) 191147.
M. Baeseler, K. Seiftarth, J. Dahlmann, A. Kunath, B. Drescher and T. Buchheim, Preparation
of alkylated phenyl alkoxy- and hydroxybenzoates,
useful as UV stabilizers, pharmaceuticals, or pesticides, via alkylation using olefins, alkyl halides, or
aryl olefins, over acidic ion-exchanger catalysts,
Ger. (East) DD 268,933 (1989); Chem. Abstr., 112
(1990) 98213.
H.J. Schmidt and H.J. Arpe, Carbonylation of
formaldehyde acetals, Eur. Pat. Appl. EP 71,920
(1983); Chem. Abstr., 99 (1983) 53151.
S.J. Lee, W.C. Lee, M.S. Lee, S.J. Oh and H.K.
Park, Condensation of methyl-N-phenylcarbamates with solid acid catalysts, Stud. Surf. Sci.
Catal., 59 (1991) 495.
R. E1 Gharbi, M. Delmas and A. Gaset, Reaction

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45

183

184

185

186

187

188

189

190
191

192

193

194

195

196

mechanism of the condensation of anethole with


acetaldehyde catalyzed by ion-exchange resins, Tetrahedron, 39 (1983)613 (Fr).
K. Okamoto, H. Kita, Y. Tanaka and S. Iimuso,
Method for preparing hisphenol A, Eur. Pat. Appl.
EP 442,122 (1991); Chem. Abstr., 115 (1991)
158713.
C.G.I. Sfintescu, I. Anton, L. Bercu, M. Bancila,
V. Iacob, T. Stan, A. Ivanovici, D. Tasalungwa
and M. Dragomir, Bisphenol production, Br. Pat.
1,410,750, (1975); Chem. Abstr., 84 (1976) 60205.
F. Heydenrich, J. Hans and N. Bachem, Catalyst
for preparation of bisphenols, US Pat. 4,369,293
(1983); Chem. Abstr., 98 (1983) 108275.
K. Jerabek, L. Beranek, K. Setinek, V. Ceska, O.
Jerabek, S. Jirasek, L. Prochazka, J. Srejber and
M. Cervinka, Bisphenols, Czech. CS 216,081
(1984); Chem. Abstr., 102 (1985).61930.
K. Jerabek, G.H. Li and K. Setinek, Comparison
of the kineics of bisphenol A synthesis on promoted and unpromoted ion exchanger catalysts,
Collect. Czech Chem. Commun., 54 (1989) 321;
Chem. Abstr., 111 (1989) 78626.
P.V. Shaw, Preparation of bisphenols by ketonephenol condensation, US Pat. 4,859,803 (1989);
Chem. Abstr., 112(1990) 36723.
Z. Jin, B. He and H. Jiang, Production technology
and apparatus of ion-exchange resin catalyst used
in bisphenol synthesis, Faming Zhuanli Shenquing
Gongkai Shoumingshu CN 1,059,480 (1992); Chem.
Abstr., 117 (1992)130921.
M.O. Nutt, Anthraquinone, US Pat. 4,304,724
(1984); Chem. Abstr., 96 (1982) 122463.
S.M. Li, Preparation of bis(p-aminocumyl)benzenes, US Pat. 4,973,754 (1990); Chem. Abstr., 114
(1991) 228500.
M.J. Astle and M.L. Pinns, Cation exchange
resin-catalyzed condensation and polymerization
of aldehydes and cyclohexanone, J. Org. Chem., 24
(1959) 56.
F. Gude and R. Winter, Process for the preparation of y-undecalactone from to-undecylenic acid,
Eur. Pat. Appl. EP 293,659 (1988); Chem. Abstr.,
111 (1989)173970.
T.F. Wood, Chemistry of Synthetic musks. II. Benzenoid musks, in fragrance chemistry, in E.T.
Theimer (Ed.), The Science and the Sense of Smell,
Academic Press, New York, 1982, p. 509.
H. Fischer, E. Karge, B. Schmidt, K. Jans, G.
Welsch, M. Wolf, R. Fiege and E. Horn, Cyanoethylation of aromatic amines, Ger. (East)DD
231,784 (1986); Chem. Abstr., 106 (1987) 1 3 9 8 2 8 .
M. Iborra, J.F. Izquiordo, J. Tejero and F. Cunill,
Ion exchange sulfonic resins as catalysts for the

39

197

198

199

200

201

202

203

204

205

206

207

208

209

decomposition of MTBE in the gaseous phase,


An. Quire., 86 (1990) 613 (Span.).
H. Reinhardt and B. Schleppinghoff, Procedure
for the preparation of tertiary olefins, Ger. Often.
DE 3,610,704 (1987); Chem. Abstr., 108 (1988)
186161.
J. Voduar, Decomposition of organic hydroperoxides on cation exchangers. IV. Kinetics of the
decomposition of p-tert-butyi-(cumene hydroperoxide), Acta Chim. Hung., 128 (1991) 861 (Eng.);
Chem. Abstr., 116 (1992) 151000.
G.S. Grover and R.V. Chaudhuri, Selective dimerization of isobutene using a cation exchange resin
catalyst, in B. Viswanathan and C.N. Pillai (Eds.),
Recent Developments in catalysis - Theory and
Practice, Narosa Publishing House, New Delhi,
1991, p. 345.
C.T. O'Connor, M. Kojima, and W.K. Schumann,
The oligomerization over cationic exchange resins,
Appl. Catal., 16 (1985) 193.
F.J. Waller, Catalysis using blends of perfluorihated ion-exchange polymers with perflorinated
diluents, US Pat. 5,105,047 (1992); Chem. Abstr.,
117 (1992) 47906.
H. Hartung, R. Hellmig, G. Heublin, H.H. Rottmayer, G. Schwachula and D. Stadermann,
Isobutene oligomerization, Ger. (East) DD 214,604
(1984); Chem. Abstr., 104 (1986) 5560.
R.M. Boden, Branched chain olefinic alcohols,
thiols, esters and ethers, organoleptic uses thereof,
and intermediates therefor, US Pat. 4,375,005
(1983); Chem. Abstr., 99 (1983)5194.
Z. Giao, B. Lu and K. Zhou, Oligomerization of
butene - cationic oligomerization catalyzed by sulfonated polystyrene resins, Shiyou Xuebao, Shiyou
Jiagong, 6 (1990) 68 (Ch.); Chem. Abstr., 114 (1991)
26108.
H.D. Koehler, Procedure for the preparation of
tetraisobutylene from diisobutylene, Ger. Often.
DE 3,612,443 (1987); Chem. Abstr., 108 (1988)
186162.
H. Hasegawa and T. Higashimura, Cationic oligomerization of styrene by solid acids. I. Catalysis by
poly(styrenesulfonic acid) resin, J. Polym. Sci.,
Polym. Chem. Ed., 18 (1980) 611.
H. Hasegawa and T. Higashimura, Cationic oligomerization of styrene by solid acids. II. Oligomerization of styrene catalyzed by perfluorinated
resinsulfonic acid (Nation-H), Polym. J., 11 (1979)
737.
A.R. Taylor, G.W. Keen and E.J. Eisenbraun,
Cyclodimerization of styrene, J. Org. Chem., 42
(1977) 3477.
B. Chaudhuri and M.M. Sharma, Some novel as-

40
pects of dimerization of a-methylstyrene with
acidic ion-exchange resins, clays and other acidic
materials as catalysts, Ind. Eng. Chem. Res., 28
(1989) 1757.
210 M.V. Enschescu, E.S. Meroiu, T.D. Bota, E.
Cerbu, Z. Cristu, L. Cringus, I. Florea and A.
Vasu, Trioxane, Rom. RO 86,013, (1985); Chem.
Abstr., 104 (1986) 129932.
211 Z. Zhang, Z. Liu and C. Guo, Synthesis of epoxidized ricebran oil, Huaxue Shijie, 31 (1990) 326
(Ch.); Chem. Abstr., 114 (1991) 165588.
212 S. Jin, Z. Zhang and Y. Jiao, Epoxidation of
cottonseed oil, Huadong Huadong Xueyuan Xuebao, 4 (1984)531 (Ch.); Chem. Abstr., 102 (1985)
202826.
213 J. Sante-Ollalo Gadea, Esters from olefin, Eur.
Pat. Appl. EP 54,576 (1982); Chem. Abstr., 97
(1982) 111605.
214 Y. Tokumoto, K. Sakamoto, K. Sasaki and I.
Shimizu, Catalyst for esterification of acetic acid
by olefins, EP 483826 (1990), Chem. Abstr., 117
(1992) 92632.
215 I.P. Stepanova, V.M. Obukhov, A.V. Bondarenko
and M.I. Farberov, Synthesis of sec-butyl acetate
by alkylation of acetic acid with n-butenes in presence of KU-23 sulfonated cation exchange resin,
Zh. Pr. Khim., 50 (1977) 640.
216 S. Yu, Catalytic synthesis of cyclohexyl carboxylate
with strong acid resin, Lizi Jiaohuan Yu Xifu, 7
(1991) 219 (Ch.): Chem. Abstr., 116 (1992) 176453.
217 A. Chakrabarti and M.M. Sharma, Cyclohexanol
from cyclohexene via cyclohexyl acetate, React.
Polym., 18 (1992)107.
218 H. Fukuhara and F. Matsunaga, Preparation of
cycloalkanols and/or their esters, Jpn. Kokai
Tokkyo Koho JP 01,254,634 (1989); Chem. Abstr.,
112 (1990) 118314.
219 S.C. Nigam, S.N. Bannore, H.N. Subbarao and
Sukh Dev, A novel process for conversion of terpenic olefins to the corresponding alcohols, ethers
of esters of perfumery value, Indian Pat. 154862
(1984); Chem. Abstr., 106 (1986) 18872.
220 L. Cerveny, A. Marhoul and J. Kozel, Addition of
acetic acid to styrene catalyzed by ion exchanger,
React. Kinet. Catal. Lett., 37 (1988) 337.
221 A. Chakrabarti and M.M. Sharma, Esterification
of acetic acid with styrene: Ion exchange resins as
catalysts, React. Polym., 16 (1991/1992) 51.
222 M.A. Perez Pascual and F. Albertas de Benito,
Preparation of secondary alkyl acrylates using
cation exchange resin catalysts, Span. ES 2,019,052,
(1991); Chem. Abstr., 115 (1991) 233105.
223 Z. Yang, W. Hu and Y. Xie, Synthesis of isobornyl
methacrylate by using ion exchange resin as catalyst, Lizi Jiaohuan Yu Xifu, 6 (1990) 178 (Ch.);
Chem. Abstr., 114 (1991)165456.

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45


224 H.W. Klein, Preparation of dicyclopentenyl esters
as perfume intermediates, Ger. Often. DE
3,619,797 (1987); Chem. Abstr., 110 (1989)153785.
225 B. Schleppinghoff, H. Leblanc and K.H. Mallmann, Carboxylic acid esters, Ger. Often. DE
3,105,399 (1982); Chem. Abstr., 98 (1983)53189.
226 A. Chakrabarti and M.M. Sharma, Anhydrous esterification of myristic acid with propylene: Ion
exchange resin and acid-treated clay as catalysts,
J. Am. Oil Chem. Soc., 69 (1992) 1251.
227 S. Okazaki, M. Kokoma, T. Kaneko and K. Sakai,
Preparation of esters, Jpn. Kokai Tokkyo Koho JP
63,297,340 (1988): Chem. Abstr., 110 (1989) 172711.
228 F. Cornea, A. Ciobanu, A.R. Nicolae and M.
Petrescu, Hexyl acetate, Rom. RO 80,597 (1982);
Chem. Abstr., 100 (1984)209186.
229 Y. Yamata, T. Murata and Y. Takagi, Acetylation
of alcohols in the presence of acidic ion exchangers, Jpn. Kokai Tokkyo Koho JP 63 30,429 (1988);
Chem. Abstr., 110 (1989) 154805.
230 Y. Yu, R. Wei, Y. Lian and R. Che, Synthesis of
glycerin carboxylate by ion exchange resin catalysts, Lizi Jiaohuan Yu Xifu, 4 (1988) 305 (Ch.);
Chem. Abstr., 111 (1989) 41752.
231 R.A. Capeletti, A.A. Chialvo, Ion exchange resins
as catalysts: Esterification of dicarboxylic acids. I.
Diethyl oxalate, Rev. Fac. Ing. Quire., 43 (1979)
155 (Span.); Chem. Abstr., 93 (1980) 138679.
232 A. Ciobanu, A. Nicolae and M. Malinetescu, Synthesis of maleic diesters, Rev. Chim., 38 (1987) 293
(Rom.); Chem. Abstr., 108 (1988) 204201.
233 H. Chi, Y. Cui and Z. Wang, Preparation of
methyl methacrylate by esterification, Faming
Zhuanli Shenqing Gongkai Shuoningshu CN
1,052,848 (1991); Chem. Abstr., 116 (1992) 84382.
234 M. Kamioka, S. Matsumoto, H. Yoshida and M.
Baba, Manufacture of methacrylic acid esters, Jpn.
Kokai Tokkyo Koho JP 03 52,842 (1991); Chem.
Abstr., 114 (1991) 247969.
235 S. Nakashimo, H. Sogabe, H. Yoshida and A.
Okubo, Process for producing unsaturated carboxylates, PCT Int. Appl. WO 88 00,180 (1988); Chem.
Abstr., 108 (1988) 205252.
236 Z. Ding and M. Wu, Synthesis of condensed ethylene glycol acrylate, Shanghai Keji Daxue Xuebao,
14 (1991) 100 (Ch.); Chem. Abstr., 116 (1992)
105596.
237 L. Jeromin, E. Peukart and G. Wollmann, Preesterification of free fatty acids in crude fats
and/or oils, Ger. Often. DE 3,501,761 (1986);
Chem. Abstr., 105 (1986) 116967.
238 M. Petrini, R. Ballini and E. Marcantoni, Amberlyst-15: A practical, mild and selective catalyst
for methyl esterification of carboxylic acids, Synth.
Commun., 18 (1988)847.
239 H. Iwasaki and M. Kobayashi, Manufacture of

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45

240

241

242

243

244

245

246

247

248

249

250

251

252

253

phenyi esters of methacrylic acid, Jpn. Kokai


Tokkyo Koho JP 63,57,554 (1988); Chem. Abstr.,
110 (1989) 8800.
T. Tono, M. Kamezawa and T. Hayashi, Manufacture of cyano group containing methacrylate esters, Jpn. Kokai Tokkyo Koho JP 02,160,753 (1990);
Chem. Abstr., 113 (1990)172924.
N. Totani and T. Muramatsu, A new procedure
for the acetylation of lipids, Chem. Phys. Lipids, 29
(1981) 375; Chem. Abstr., 96 (1982)103596.
M.L. Wilson and J.C. Chen, Monoesterification of
substituted pyrrole diacids, Ger. Often. 2,942,186
(1980); Chem. Abstr., 93 (1980) 150114.
A. Rataj., Kulicki and A. Krogulocki, Process for
producing glycol monoether acetates, Pol. PL
137,483 (1987); Chem. Abstr., 114 (1991) 45464.
J.L. Platt Jr., Catalytic acylation of dimethyl phosphoramidothioate with anhydrides, US Pat.
3,732,344 (1973); Chem. Abstr., 79 (1973) 31481.
Mitsubishi Chemical Industries Co., Ltd., Glycolic
acid ethers and estes, Jpn. Kokai Tokkyo Koho JP
82 28,015 (1982); Chem. Abstr., 96 (1982) 217254.
K. Setinek, Z. Prokop, M. Kraus and Z. Zitny, Ion
exchanger-catalyzed addition of alkenes to alcohols, Chem. Prum., 27 (1977) 451 (Czech.); Chem.
Abstr., 88 (1978)89028.
J.D. Chase, B.B. Galvez and H.J. Woods, Etherification of light hydrocarbons, US Pat. 4,204,077
(1980); Chem. Abstr., 93 (1980) 149793.
J. Van Pool, Aliphatic ether production, Eur. Pat.
Appl. EP 47,906 (1982); Chem. Abstr., 96 (1982)
219662.
S.T. Lee and W.Y. Lee, A study on the liquid
phase reaction rate of methyl-tert-butyl ether
preparation on ion exchange resin catalyst, Kongdae Yon'gu Pogo, 13 (1981) 7 (Korean); Chem.
Abstr., 98 (1983) 71181.
F. Colombo and L. Dolloro, MTBE synthesis:
Activity of different catalysts and kinetics, Geterog.
Katal., 5 (1983) 407; Chem. Abstr., 100 (1984)
208782.
W.G. Rothschild, Ethers and alcohols, Eur. Pat.
Appl. EP 102,840 (1984); Chem. Abstr., 101 (1984)
90405.
A.M. Jarallah, M.A.B. Siddiquih and A.K.K. Lee,
Kinetics of methyl tert-butyl ether synthesis catalyzed by ion exchange resin, Can. J. Chem. Eng.,
66 (1988) 802.
J.M. Adams, K. Martin, R.W. McCabe and S.
Murray, Methyl-tert-butyl ether (MTBE) production: A comparison of montmorillonite-derived
catalysts with an ion-exchange resin, Clays Clay
Miner., 34 (1986) 597.

41
254 J. Tejero, F. Cunill and M. Iborra, Molecular
mechanisms of MTBE synthesis on a sulfonic acid
ion exchange resin, J. Mol. Catal., 42 (1987) 257.
255 P.P. Kapustin, S.T. Guliyants, A.P. Vorozheikin,
Yu. I. Ryazanov, P.A. Kirpichnikov, A.G. Liakumovich, N.L. Kozhin, N.G. Cherkasov, R.A.
Akhmed'yanova, et al., Method of producing
methyl C a- or Cs-tert-alkyl ethers, U.S.S.R. SU
1,527,233 (1989); Chem. Abstr., 112 (1990) 197623.
256 P.P. Kapustin, S.T. Guliyants, A.P. Vorozheikin,
Yu. I. Ryazanov, P.A. Kirpichnikov, A.G. Liakumovich, N.L. Kozhin, N.G. Cherkasov, R.A.
Akhmed'yanova, et al., Method of producing
methyl C a- or Cs-tert-alkyl ethers, U.S.S.R. SU
1,527,232 (1989); Chem. Abstr., 112 (1990) 197624.
257 A. Hejtmankova, K. Jerabek and K. Setinek, Kinetics of methyl-tert-butyl esther synthesis in
gaseous phase, Collect. Czech. Chem. Commun., 55
(1990) 1033; Chem. Abstr., 113 (1990) 77423.
258 A.G. Liakumovich, S.T. Guliyants, P.P. Kapustin,
P.A. Kirpichnikov, A.P. Vorozheikin, Yu. I.
Ryazanov, R.A. Akhmed'yanova, N.L. Kozhin,
V.A. Kurbatov, et al., Method of obtaining
methyl-tert-alkyl C a or C 5 ethers, U.S.S.R. SU
1,565,835 (1990); Chem. Abstr., 113 (1990) 171495.
259 D.K. Lee, S.B. Kim, B. Sung, and J.H. Ahn,
Changes in characteristics of MTBE synthesis and
cracking with sulfonic acid group distributions in
porous poly(styrene-co-divinyl-benzene) resin catalysts, Hwahak Konghak, 29(1) (1991) 49; Chem.
Abstr., 115 (1991) 91626.
260 A. Hejtmankova, K. Jerabek and K. Setinek, Kinetics of methyl tert-butyl ether synthesis in
gaseous phase, Collect. Czech. Chem. Commun., 55
(1990) 1033.
261 G.J. Hutchings, C.P. Nicolaides and M.S. Scurrell,
Developments in the production of methyl tertbutyl ether, Catal. Today, 15 (1992)23.
262 Li-Man Tau and B.H. Davis, Acid catalyzed formation of ethyl tertiary butyl ether (ETBE), Appl.
Catal., 53 (1989) 263.
263 O. Francoisse and F.C. Thyrion, Kinetics and
mechanism of ethyl tert-butyl ether liquid-phase
synthesis, Chem. Eng. Process., 30 (1991) 141.
264 S. Randriamahefa, R. Gallo, G. Raoult and P.
Mulard, Synthese de L'ether methyl ter-amylique
(TAME) en catalyse acide: Cinetiques et equilibres de la methoxylation du methyl-2-butene-2, J.
Mol. Catal., 49 (1988)85.
265 A.O.I. Krause and L.G. Hammarstr6m, Etherification of isoamylenes with methanol, Appl. Catal.,
30 (1987)313.
266 I. Muja, D. Goidea, N. Marculescu, G. Andreescu,

42
C.I. Curca and I. Antonescu, Aditia metanolului
la izoamilene pe cationiti macroporosi, Evista De
Chimie, 37(12)(1986)1047.
267 S.S. Jayadeokar and M.M. Sharma, Simultaneous
hydration and etherification of isoamylene using
sub-azeotropic ethanol: Use of ion exchange resins
as catalyst, React. Polym., 19 (1993) 169.
268 J. Wawak, K. Fraczek, S. Manka, M. Konopka and
Z. Waligorski, Manufacture of ethylene glycol
mono-tert-butyl ether for coating solvent, Pol. PL
153,192 (1991); Chem. Abstr., 116 (1992) 61634.
269 J. Wawak, K. Fraczek, S. Manka, M. Konopka and
Z. Waligorski, Manufacture of ethylene glycol
mono-tert-butyl ether for coating solvent, Pol. PL
153,194, (1991); Chem. Abstr., 116 (1992) 61635.
270 M. Imaizumi and M. Yasuda, Glycol ethers, Ger.
Often. DE 2,450,667 (1975); Chem. Abstr., 83
(1975) 178305.
271 O. Matsumoto, T. Naguchi and H. Fujiwara, Preparation of propylene glycol tert-butyl ether, Jpn.
Kokai Tokkyo Koho JP 63,250,336 (1988); Chem.
Abstr., 110 (1989)114304.
272 K. Nyi and A.W. Kohr, Monoether addition: Product of the reaction between a multivalent compound with dicyclopentadiene, Ger. Often.
2,833,597 (1979); Chem. Abstr., 90 (1979) 203563.
273 S.C. Nigam, S.N. Bannore, H.N. Subbarao and
Sukh Dev, A Novel process for conversion of
terpenic olefins to the corresponding alcohols,
ethers or esters of perfumery value, Indian Pat.
154861; Chem. Abstr., 106 (1986) 5268.
274 J.E. Lyons and P. Hosler, Ethers by catalytic reaction of an olefin with an alcohol, Ger. Often.
2,706,879 (1977); Chem. Abstr., 87 (1977) 151700.
275 S.S. Jayadeokar and M.M. Sharma, unpublished
Work.
276 J. Kiwala, R.J. Takarzewski, F.L. Schmitt and M.A.
Sprecker, Phenethyl cyclohexyl ethers, US Pat.
4,324,923 (1982); Chem. Abstr., 97 (1982) 72068.
277 A. Chakrabarti and M.M. Sharma, unpublished
work.
278 A. Chakrabarti and M.M. Sharma, unpublished
work.
279 Daicel Chemical Industries Ltd., Manufacture of
glycosides, Jpn. Kokai Tokkyo Koho JP 60 01,196
(1985); Chem. Abstr., 102 (1985) 221145.
280 R. Lebuhn, J. Feldmann and H. Koebernick,
Preparation of color-stable methylglucoside, Ger.
Often. DE 3,611,035 (1987); Chem. Abstr., 108
(1988) 112891.
281 A. Meissler, H. Winter, R. Neumann and M.
Fedtke, Direct production of methylolurea etherifled with higher alcohols, Ger. (East) DD 220,954
(1985); Chem. Abstr., 104 (1986)109055.
282 W. Neier and J. Woellner, Isopropyl alcohol by
direct hydration, Chem. Tech, 3 (1973) 95.

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45


283 E. Velo, L. Puigjaner and F. Recasens, Intraparticle mass transfer in the liquid-phase hydration of
isobutene: Effects of liquid viscosity and excess
product, Ind. Eng. Chem. Res., 29 (1990) 1485.
284 A. Delion, B. Torck and M. Hellin, Hydration of
isopentenes in an acetone environment over ionexchange resin: Thermodynamic and kinetic analysis, J. Catal., 103 (1987) 177.
285 Z. Prokop and K. Setinek, Hydration of 2-methylbutenes on organic ion exchange resin catalysts,
Collect. Czech. Chem. Commun., 52 (1987) 1272.
286 B. Schleppinghoff, R. Malessa, C. Gabel, H.V.
Scheef and M. Lux, Preparation of tertiary amyl
alcohol by hydration of isoamylenes on acidic
cation exchangers, Ger. Often. DE 3,801,273
(1989); Chem. Abstr., 112 (1990) 7031.
287 R. Malessa and B. Schleppinghoff, Preparation of
C4_5 tertiary alcohols by hydration of C4.5 tertiary
olefins on acidic cation exchangers, Ger. Often.
DE 3,801,275 (1989); Chem. Abstr., 112 (1990)
7032.
288 S.C. Nigam, S.N. Bannore, H.N. Subbarao and
Sukh Dev, A novel process for conversion of terpinic olefins to the corresponding alcohols, ethers
or esters of perfumery value, Indian Pat. 154860
(1984); Chem. Abstr., 105 (1986) 227075.
289 B. He and Y. Liu, Preparation of camphor via
hydration of camphene in presence of ion exchange resins, Faming Zhuanli Shenqing Gongkai
Shuomingshu CN 85,104,105 (1986); Chem. Abstr.,
109 (1988) 23144.
290 Q. Chen, X. Cai and S. Cao, et al., Preparation of
isoborneol by resin-catalyzed hydration of camphene, Faming Zhuanli Shenqing Gongkai Shuomingshu CN 1,049,842 (1991); Chem. Abstr., 115
(1991) 159483.
291 L.B. Wasker and V.G. Pangarkar, Catalysis with
ion exchange resins: Hydration of citronellol, Indian Chem. Eng., 34(2/3) (1992) 99.
292 D. Arntz and N. Wiegand, Preparation of 1,3-propanediol by hydration of acrolein over phosphonate or aminophosphonate group containing ion
exchange resins and subsequent catalytic hydrogenation, Eur. Pat. Appl. EP 412,337 (1991); Chem.
Abstr., 114 (1991) 142663.
293 D. Arntz and N. Wiegand, Process for the preparation of 1,3-propanediol via hydration of acrolein
using chelate-forming ion exchangers as catalysts,
Eur. Pat. 487,903 (1992); Chem. Abstr., 117 (1992)
111131.
294 a. K. Matsuoka and M. Yamada, Preparation of
1,2-cyclohexanediol, Jpn. Kokai Tokkyo Koho JP
03,236,337 (1991); Chem. Abstr., 116 (1992) 128235.
294 b. K. Riedel and H. Krekeler, 3-Hydroxycyclohexanone, Ger. Often., 2,205,225 (1973); Chem.
Abstr., 79 (1973) 115215.

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45


295 J. Sun and Q. Wang, Copper catalyst for hydration
of nitrile groups, Faming Zhuanli Shenqing
Gongkai Shuomingshu CN 1,044,417 (1990); Chem.
Abstr., 116 (1992)47145.
296 Y. Fuchigami, Hydrolysis of methyl acetate in distillation column packed with reactive packing of
ion exchange resin, J. Chem. Eng. Jpn., 23 (1990)
354.
297 E. Drent, Preparation of alcohols from olefins,
Brit. UK Pat. Appl. GB 2,238,539 (1991); Chem.
Abstr., 115 (1991) 207509.
298 Maruzne Oil Co. Ltd., Hydrolysis of mono-tertiary-alkyl ethers of alkene glycols, Jpn. Kokai
Tokkyo Koho JP 81,138,125 (1981); Chem. Abstr.,
96 (1982) 85044.
299 T. Kinya, K. Hiroki, N. Shigenori, K. Takashi, W.
Tadahiro and M. Tadashi, Hydrolysis of glycol
dialkyl ether to a monoalkyl ether and tertiary
alcohol or tertiary olefin, Eur. Pat. Appl. EP28,129
(1981); Chem. Abstr., 96 (1982) 85047.
300 J. Van Gogh, A two-step monoethylene glycol
preparation process, Eur. Pat. Appl. EP 448,157
(1991); Chem. Abstr., 115 (1991) 207515.
301 K. Ataka, K. Omori and M. Takahashi, Process
for the preparation of monosodium phosphoenolpyruvate monohydrate as blood preservatives,
Jpn. Kokai Tokkyo Koho JP 63 45,288 (1988);
Chem. Abstr., 109 (1988) 148901.
302 T. Nakamura, K. Yoshido and K. Nakada, Isomerization of dicyclopentadiene, Jpn. Kokai Tokkyo
Koho 02 17,130 (1990); Chem. Abstr., 112 (1990)
199334.
303 R.C. Michaelson and G. Cerri, Isomerization of
2-methyl-l-butene to 2-methyl-2-butene in an
isoamylene mixture, PCT Int. Appl. WO 89 02,882
(1989); Chem. Abstr., 111 (1989) 40086.
304 J.H. Ahm, S.K. Ihm and K.S. Park, Effects of the
local concentration and distribution of sulfonic
acid groups on 1-butene isomerization catlayzed
by macroporous ion-exchange resin catalysts, J.
Catal., 113 (1988) 434.
305 K. Takahashi, T. Nakayama and Y. Tao, Preparation of 2,5-dimethyl-2,4-hexadiene by isomerization, Jpn. Kokai Tokkyo Koho JP 03,232,822
(1991); Chem. Abstr., 116 (1992) 105604.
306 Toyo Soda Mfg. Co., Ltd., Phenylacetaldehyde,
Jpn. Kokai Tokkyo Koho JP 82 18,643 (1982);
Chem. Abstr., 97 (1982) 98670.
307 D.J. Casey and W.J. Trzaskos, Synthesis of 10-hydroxymethyl-9-anthraldehyde, US Pat. 4,593,130
(1986); Chem. Abstr., 105 (1986)133547.
308 S.M.K. Li, Process and catalysts for the isomerization of bisphenol A byproduct isomers to bisphenol A, US Pat. 4,822,923 (1989); Chem. Abstr., 111
(1989) 78788.
309 S.M.K. Li, A process for isomerizing the byprod-

43
ucts of the bisphenol synthesis, Eur. Pat. Appl. EP
313,165 (1989); Chem. Abstr., 112(1990) 7156.
310 Y. Natsuume, Liquid dibasic acid anhydrides, Jpn.
Kokai Tokkyo Koho JP 78,130,638 (1978); Chem.
Abstr., 90 (1979) 88340.
311 W. Swodenk, W. Schwerdtel and P. Losacker,
Synthesis of isoprene from isobutylene and formaldehyde, Erdoel Kohle, Erdgas, Petrochem, 23
(1970) 641 (Ger.); Chem. Abstr., 74 (1971) 4389.
312 M. Delmas, P. Kalck, J.P. Gorrichon and A. Gaset,
An easy studient synthesis of a substituted 1,3-dioxane by use of an ion exchange resin as catalyst:
Clean illustration of the Prins reaction, J. Chem.
Educ., 59 (1982) 700.
313 M. Deimas and A. Gaset, Supported acid catlaysis
with ion-exchange resins. III. Condensation between styrene and polymerized formaldehyde in
organic media, J. Mol. Catal., 17 (1982)35.
314 J.C. Jain, M.K. Sahni, K.C. Gupta and C.K.
Narang, Some resin catalyzed synthetic reactions
of amino acids and peptides, in G.T. Gadre (Ed.),
Proc. lon-Exch. Symp., 1978, Central Salt Mar.
Chem. Res. Inst., Bhavnagar, India; Chem. Abstr.,
92 (1980)6886.
315 E. Gutierrez, A.J. Aznar and E. Ruiz-Hitzky, Selectivity of the catalytic rearrangement of 1,20-glycols on acidic solids, Stud. Surf. Sci. Catal., 41
(1988) 211; Chem. Abstr., 110 (1989) 192023.
316 a. K.G. Devenport and C.B. Hilton, 4-Acetoxyacetanilide, US Pat. 4,560,789 (1985); Chem. Abstr.,
104 (1986)148552.
316 b. K.R. Lassila and M.E. Ford, Solid acid catalysis
of the Fries rearrangement: Thermodynamic limitations based on solvent polarity, in W.E. Pascoe
(Ed.), Catalysis of Organic Reactions, Marcel
Dekker, New York, 1992, p. 169.
317 N. Kalyanam and S. Sivaram, Cation exchange
resin-catalyzed Ritter reaction of isobutylene with
acetonitrile: A remarkable role of bound water, in
D. Naden and M. Streat (Eds.), Ion Exchange
Technology, Ellis Horwood, Chichester, 1984, p.
458.
318 R.B. Perni, Amberlyst-15 as a convenient catalyst
for chemoselective thioacetalization, Synth. Commun., 19 (1989) 2383.
319 G.A. Olah, G.K.S. Prakash, P.S. Iyer, M. Tashiro
and T. Yamato, Nafion-H catalyzed de-tert-butylation of aromatic compounds, J. Org. Chem., 52
(1987) 1881.
320 T. Yamato, C. Hideshima, A. Miyazawa, M.
Tashiro, G.K.S. Prakash and G.A. Olah, Solid
superacids catalyzed organic synthesis. 2. Nafion-H
catalyzed tert-butylation of aromatic compounds
with 2,6-di-tert-butyl-p-cresol, Catal. Lett., 6 (345)
1990.
321 R.G. Duranleau, E.C.Y. Nieh and J.F. Knifton,

44
Process for production of ethylene glycol and
dimethyl carbonate, US Pat. 4,691,041 (1987);
Chem. Abstr., 107 (1987)200935.
322 M.J. Mullins and P.J. Hamlin, Allylic compounds,
US Pat. 4,590,300 (1986); Chem. Abstr., 105 (1986)
97004.
323 J. Disteldorf, W. Huebel and L. Broschinski, Primary mono- and diamines from oxo compounds,
Ger. Often. DE 3,021,955 (1981); Chem. Abstr., 96
(1982) 85124.
324 S. Doi, Y. Takayanagi and Y. Hara, Method of
making N-alkoxymethyl(meth)acrylamides, Jpn.
Kokai Tokkyo Koho JP 62,281,849 (1987); Chem.
Abstr., 109 (1988)38405.
325 S. Doi and Y. Takayanagi, Manufacture of N-substituted cyclic imides from dicarboxylic anhydrides
and amines, Jpn. Kokai Tokkyo Koho JP
62,212,361 (1987); Chem. Abstr., 109 (1988) 54642.
326 S. Doi and Y. Takayanagi, Manufacture of
bis(cyclic imides) from dicarboxylic anhydrides and
diamines, Jpn. Kokai Tokkyo Koho JP 62,234,062
(1987); Chem. Abstr., 109 (1988) 92751.
327 S. Doi and Y. Takayanagi, Ion exchange resin
catalysts for bismaleimide manufacture, Jpn. Kokai
Tokkyo Koho JP 62,221,668 (1987); Chem. Abstr.,
108 (1988) 132456.
328 F. Ladhar, R. E1Gharbi, M. Delmas and A. Gaset,
Synthesis of 1,3,5-triacylperhydro-l,3,5-triazines
catalyzed by ion exchange resins, Chem. Abstr.,
107 (1987) 77759.
329 A. Balogh, L. Kukacka, J. Durmis, M. Karvas, P.
Moravec, J. Masek, J. Sabados, M. Collak, M.
Magura and M. Holko, 2,2,6,6-Tetramethyl-4piperidinone, Czech CS 238,851 (1987); Chem. Absir., 107 (1987) 178627.
330 J. Andrade and G. Prescher, Preparation of 2,5-dimethoxy- and 2,5-diethyoxytetrahydrofuran, Ger.
Often. DE 3,544,046 (1987); Chem. Abstr., 107
(1987) 198068.
331 J. Kampe and G. Kiessling, Preparation of 2,2dimethyl-4-(hydroxymethyl)-l,3-dioxolane, Ger.
(East) DD 238,233 (1986); Chem. Abstr., 106 (1987)
67287.
332 Y. Shimizu and M. Fukuda, Unsaturated cornpounds having 1,3-dioxolane rings, Jpn. Kokai
Tokkyo Koho JP 60,208,974 (1985); Chem. Abstr.,
104 (1986) 168451.
333 N. Fujishige, R. Numajiri and M. Tegi, Hydroxylated xanthan gums, Jpn. Kokai Tokkyo Koho JP
60,252,601 (1985); Chem. Abstr., 105 (1986) 99460.
334 R. Fischer and J. Paust, 2-Alkyl-4,4-diacyloxy-2butenals, Ger. Often. DE 3,418,747 (1985); Chem.
Abstr., 104 (1986) 167996.
335 Nittetsu Chemical Industrial Co., Liquid hydrocarbon resins and their preparation, Jpn. Kokai

A. Chakrabarti, M.M. Sharma / React. Polym. 20 (1993) 1-45

336

337

338

339

340

341

342

343

344

345

346

347

348

349

350

351

Tokkyo Koho JP 59,108,014 (1984); Chem. Abstr.,


102 (1985)25557.
R.S. Nevin, Lactic acid-glycolic acid copolymers,
Fr. Pat. 2,464,973 (1981); Chem. Abstr., 96 (1982)
35892.
J. Feldmann, H. Koebernik and H.U. Woelk, Anhydro polyol containing polyol mixtures, Can. CA
1,178,288 (1984); Chem. Abstr., 102 (1985) 185434.
A. Denis, M. Delmas and A. Gaset, Dioxin heterocycles. New method for the synthesis of 4Hbenzo-l,3-dioxins, J. Heterocycl. Chem., 21 (1984)
517 (Fr); Chem. Abstr., 101 (1984) 171187.
M.V. Enachescu, E.S. Meroiu, T.D. Bota, E.
Cerbu, Z. Cristu, L. Cringus, I. Florea and A.
Vasu, Cyclotrimerization of formaldehyde, Rom.
RO 86,013 (1985); Chem. Abstr., 104 (1986) 129932.
W.J. Greenlee and E.D. Thorsett, Mild conversion
of carboxamides and carboxylic acid hydrazides to
acids, J. Org. Chem.,46(1981) 5351.
Chisso Corporation, Composition for silica coatings, Jpn. Kokai Tokkyo Koho JP 82 94,057 (1982);
Chem. Abstr., 97 (1982) 164690.
H.D. Krall and H.H. Schwarz, Azomethines, Ger.
Often. 2,525,295 (1976); Chem. Abstr., 86 (1977)
89367.
J.E. Corn and J.L. Webb, Tetrahydrofuran, Can.
CA 1,037,485 (1978); Chem. Abstr., 90 (1979)
103804.
M.L. Listemann, R. Pierantozzi and R.K. Pinschmidt, Preparation of secondary formamides by
reaction of acetals with farmamide, Eur. Pat. Appl.
EP 332,083 (1989); Chem. Abstr., 112(1990)98034.
E. Arretz, Process for the preparation of secondary mercaptans from linear olefins, Eur. Pat.
Appl. EP 329,520 (1989); Chem. Abstr., 112 (1990)
58743.
U. Rudolph and C. Wulff, Disproportionation of
bisphenols, Ger. Often. DE 3,827,643 (1990);
Chem. Abstr., 113 (1990) 61717.
J. Sun and Q. Wang, Phenolic resin synthesis
catalyzed by cation exchange resin, Faming
Zhuanli Shenqing Gongkai Shuomingshu CN
1,045,402 (1990); Chem. Abstr., 114 (1991) 165101.
F.J. Waller, Preparation of N-substituted carbamates, US Pat. 4,987,248 (1991); Chem. Abstr., 114
(1991) 184825.
B.A.O. Alink, Synthesis of aryl-substituted aliphatic acids, US Pat. 5,034,161 (1991); Chem. Abstr.,
115 (1991) 231867.
A.K. Maskaev, E.N. Pavlikova, M.V. Kotlova, S.G.
Yazepova and I.F. Kobilinskaya, Acetylation of
oleic acid in the presence of cation exchange resins,
Maslo-Zhir. Prom-st., 2 (1977) 26 (Russ.); Chem.
Abstr., 86 (1977) 157396.
S.M. EI-Shamio, F. Bassyoumi and M.H. El-mal-

A. Chakrabarti, M.M. Sharma /React. Polym. 20 (1993) 1-45


lah, Acetoxylation (hydroxylation) of oleic acid,
Grasas Acetites (Seville), 40 (1989) 272; Chem.
Abstr., 115 (1991) 70885.
352 T. Yamato, N. Sakane, T. Furusawa, M. Tashiro,
G.K.S. Prakash and G.A. Olah, Organic reactions
catalyzed by solid superacids. Part 9. FriedelCrafts cyclobenzylation of benzene with bis(hydroxymethyi)diarylalkanes catalyzed by a perfluorinated sulfonic acid resin (Nation H), J. Chem.
Res., Synop., 9 (1991) 242; Chem. Abstr., 115 (1991)
255765.
353 R.A. Bunce, R.E. Drumright and V.L. Taylor,
Amberlyst-15 promoted synthesis of fused and
spiro y-butyrolactones from /3-hydroxyesters,
Synth. Commun., 19 (1989) 2423.
354 S. Hu, Preparation of phenolphthalein using sulfonic acid-type cation exchange resin ans catalyst,
Lizi Jiaohuan Yu Xifu, 5 (1989) 454 (Ch); Chem.
Abstr., 114 (1991)61865.
355 J.P. Schirmann and M. Pralus, Process for manufacturing hydrogen peroxide and isobutene, Eur.
Pat. Appl. EP 237,402 (1987); Chem. Abstr., 108
(1988) 37215.
356 T. Yamato, C. Hideshima, G.K.S. Prakash and
G.A. Olah, Solid superacid-catalyzed organic synthesis. 4. Perflurinated resinsulfonic acid (NationH) catalyzed Friedel-Crafts benzylation of benzene and substituted benzemes, J. Org. Chem., 56
(1991) 2089.
357 T. Yamato, C. Hideshima, M. Tashiro, G.K.S.
Prakash and G.A. Olah, Solid superacids catalyzed organic synthesis. 3. Nation-H catlayzed
condensation of acetophenone derivatives. A
preparative route to 1,3,5-triarylbenzenes, Catal.
Lett., 6(1990) 341.
358 T. Yamato, C. Hideshima, G.K.S. Prakash and
G.A. Olah, Solid superacid catalyzed organic synthesis. 6. Perfluorinated resinsulfonic acid (Nafion-H) catalyzed ring closure reaction of 2,2'-dihydroxybiphenyls. A preparative route to dibenzofurans, J. Org. Chem., 56 (1991) 3192.
359 N.F. Shah and M.M. Sharma, unpublished Work.
360 A.S. Podberezina, S.T. Klimakhina, N.A. Novikov
and L.A. Kheifits, Acylation of some branched
alkenes in the presence of a KU-23 cation exchanger, Zh. Vses. Khim. O-va., 33 (1988) 350
(Russ); Chem. Abstr., 110 (1989) 172666.

45
361 A.S. Podberezina, A.S., S.T. Klimakhina, N.d.
Sergeeva, and L.A. Kheitits, Acetylation of diisobutylene by acetic anhydride in the presence of
a cation exchanger, Zh. Vses. Khim. O-va., 23
(1978) 343; Chem. Abstr., 89 (1978) 117529.
362 N.F. Shah and M.M. Sharma, A covenient method
for the preparation of methyl ketones: acetylation
of diisobutylene and diisoamylene with acetic anhydride in the presence of cation exchange resin
and acid-treated clay as catalysts, React. Polym.,
20 (1993) 47.
363 L.B. Wasker, V.G. Pangarkar and S.G. Patnekar,
Amberlyst-15 catalyzed cyclization of citronellal,
unpublished work.
364 Y. Zhang and M. Zhang, Catalytic hydrogenation
of fatty oil with ion exchange resin supported
palladium complexes, Huaxue Shiji, 10 (1988) 121
(Ch); Chem. Abstr., 110 (1989) 77999.
365 K. Setinek, S. Drapakova and Z. Prokop, Oxidation of cumene by molecular oxygen on heterogenized cobalt catalysts, Collect. Czech. Chem. Cornmun., 51 (1986) 1958; Chem. Abstr., 107 (1987)
6622.
366 V.D. Kopylova, E.L. Frumkina, L.A. Mochalova
and K.M. Saldadze, Study of the catalytic activity
of ion exchange complexes in the decomposition
of hydrogen peroxide, Deposited Document,
(1980), VINITI 3327-80; Chem. Abstr., 96 (1982)
12055.
367 R.A. Sheldon, Heterogeneous catalytic oxidation
and tinechemicals, Stud. Surf Sci. Catal., 59(1991)
33.
368 R. Ishikawa, H. Tsuneki, H. Sugizawa, Y. Shimasaki, T. Hayashi, T. Hayashi and M. Ueshima,
Process for producing alkylenamines from
aziridines and amines, Eur. Pat. Appi. EP 375,279
(1990); Chem. Abstr., 114 (1991) 81020.
369 J.M. Campbell, Acidity and Properties of Major
IndustrialAcids, Chapman and Hall, London, 1992.
370 P. Rys and W.J. Steinegger, Acidity functions of
solid-bound acids, J. Am. Chem. Soc., 101 (1979)
4801.
371 G.A. Olah, G.K.S. Praksh, and J. Sommer, Superacids, Science, 206 (1979)13.
372 P. Laszio, Catalysis of organic reactions by inorganic solids, Acc. Chem. Res., 19 (1986) 121.

You might also like