You are on page 1of 13

See

discussions, stats, and author profiles for this publication at:


https://www.researchgate.net/publication/228901786

Alkylation of aromatics with ethylene and


propylene: recent developments in
commercial processes
Article in Applied Catalysis A General November 2001
DOI: 10.1016/S0926-860X(01)00807-9

CITATIONS

READS

130

387

3 authors, including:
Thomas Degnan
University of Notre Dame
59 PUBLICATIONS 919 CITATIONS
SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

Available from: Thomas Degnan


Retrieved on: 13 September 2016

Applied Catalysis A: General 221 (2001) 283294

Alkylation of aromatics with ethylene and propylene: recent


developments in commercial processes
Thomas F. Degnan Jr. , C. Morris Smith, Chaya R. Venkat
ExxonMobil Research and Engineering Company, Annandale, NJ 08801, USA

Abstract
This paper provides an overview of current industrial alkylation processes for the production of ethylbenzene and cumene.
In recent years, zeolite catalysts have begun to displace the conventional aluminum chloride and solid phosphoric acid (SPA)
FriedelCrafts catalysts used in both ethylbenzene and cumene processes. This transformation has been particularly rapid in
the case of cumene technology, where more than 50% of the worldwide cumene capacity has converted to zeolite catalysts in
the last 5 years. 2001 Elsevier Science B.V. All rights reserved.
Keywords: Industrial processes; Cumene; Ethylbenzene; Aromatics alkylation; Zeolites

1. Introduction
Ethylbenzene and cumene are commercially the
two largest volume derivatives of benzene. Combined,
these two chemicals account for nearly 75% of the
worlds consumption of petrochemical grade benzene.
The markets and the applications for both chemicals
are fairly well established. The worldwide capacity for
ethylbenzene is currently estimated to be 23 million
metric tons per year (1 metric ton = 1000 kg) with an
annual growth rate projected to be approximately 4%.
Over 90% of the worlds production of ethylbenzene
is used in the manufacture of styrene. Other applications include paint solvents and pharmaceuticals.
Cumene capacity topped 9.5 million metric tons in
1998 and is projected to reach 10.4 million metric tons
by the end of 2003 [1]. Like ethylbenzene, cumene is
used almost exclusively as a chemical intermediate.
Its primary use is in the co-production of phenol and
Corresponding author. Tel.: +1-609-224-2875.
E-mail address: tfdegnan@pau.mobil.com (T.F. Degnan Jr.).

acetone through cumene peroxidation. Phenolic resins


and bisphenol A are the main end uses for phenol.
Bisphenol A, which is produced from phenol and acetone, has been the main driving force behind increased
phenol demand. Its end use applications are in polycarbonate and epoxy resins. The growth rate of cumene
is closely related to that of phenol and is expected to
be approximately 5.1% per year worldwide over the
next 5 years. Process technologies for both chemicals
have been moving away from conventional aluminum
chloride and phosphoric acid catalyzed FriedelCrafts
alkylation of benzene, toward zeolite-based processes.
1.1. Ethylbenzene processes
Nearly 40% of the worlds ethylbenzene capacity
still utilizes the AlCl3 -based FriedelCrafts alkylation
process first introduced in the 1950s. This process
operates at low benzene:ethylene (B:E) ratio (2:3.5)
and reasonable temperatures (250 C). However, over
the past 20 years, the operating costs associated
with the corrosivity of AlCl3 , and other problems

0926-860X/01/$ see front matter 2001 Elsevier Science B.V. All rights reserved.
PII: S 0 9 2 6 - 8 6 0 X ( 0 1 ) 0 0 8 0 7 - 9

284

T.F. Degnan Jr. et al. / Applied Catalysis A: General 221 (2001) 283294

Table 1
Commercial zeolite-based ethylbenzene processes
Process

Year first
announced

Licensed capacity
(million metric
ton per year)

MobilBadger ethylbenzene
MobilBadger dilute ethylene
MobilBadger EBMax
LummusUOP
SINOPEC
CDTech
Dow Chemical
dilute ethylene
Albene

1976
1987
1995
1989
1993
1995
1998

8350
820
5000
3650
30
880
Unknown

1992

Unknown

associated with its safe handling and disposal have


induced most manufacturers to move toward zeolite
catalyzed processes. Zeolite catalyzed processes are
licensed by MobilBadger, LummusUOP, CDTech,
and Dow Chemical. Table 1 summarizes the manufacturing volume share of each of the licensed process
technologies for ethylbenzene manufacture. Over the
last several years, the industry has moved to more
selective ethylbenzene dehydrogenation catalysts,
making the purity of ethylbenzene of even greater importance. Fig. 1 shows the major reactions associated

with ethylbenzene synthesis. Fig. 2 shows the major


side reactions. The selectivity to the monoethylated
product is very important in ethylbenzene synthesis, whereas ethylene oligomerization is normally a
secondary consideration.
1.2. MobilBadger ethylbenzene process
First introduced in 1980, the MobilBadger vapor
phase process is still the most widely used zeolite
catalyzed ethylbenzene manufacturing process. Since
its initial commercial application in 1981, over 35
units have been licensed with a total annual capacity
of nearly 8 million metric tons. In the MobilBadger
process, benzene is alkylated in a fixed-bed reactor with ethylene in the gas phase using a ZSM-5
based catalyst [2,3]. Fig. 3 is a simplified process
flow diagram of the first generation MobilBadger
process. Fresh and recycled benzene is combined
with a diethylbenzene-rich recycle stream and fed,
together with fresh ethylene, to an alkylation reactor.
The process operates with a B:E ratio of 5:20 (M) at
temperatures ranging from 370 to 420 C and pressures ranging from 0.69 to 2.76 MPa (6.827.2 bar).
Weight hourly space velocities (WHSVs), based on

Fig. 1. Primary alkylation and transalkylation chemistry for ethylbenzene production.

T.F. Degnan Jr. et al. / Applied Catalysis A: General 221 (2001) 283294

Fig. 2. Side reactions typical for the acid catalyzed reaction of benzene with ethylene.

Fig. 3. The MobilBadger vapor-phase ethylbenzene process.

285

286

T.F. Degnan Jr. et al. / Applied Catalysis A: General 221 (2001) 283294

fresh feed, range from 300 to 400 hr1 . The process


has an overall ethylbenzene yield of at least 99.0%.
Together with the Badger Technology Center of
Raytheon Engineers & Constructors, Mobil introduced second and third generation processes in 1986
and 1991. These next generation processes provided
improved cycle lengths and higher yields [4]. Cycle
lengths improved from 60 days to more than 1 year,
and yields improved to greater than 99.5%. Selectivity losses to both heavy aromatics and xylenes
dropped by more than 50%. The energy efficiency
of the MobilBadger process is reported to be high.
Essentially, all of the exothermic heat of reaction is
recovered.
The MobilBadger process has also been modified
to use dilute ethylene from FCC off-gas streams as a
feedstock. The dilute ethylene process was first commercialized at Shells Stanlow UK refinery in 1991
and has been in continuous use ever since.
1.3. MobilBadger EBMax process
EBMax is a liquid phase ethylbenzene process that
uses Mobils proprietary MCM-22 zeolite as the catalyst. This process was first commercialized at the
Chiba Styrene Monomer Co., Chiba, Japan in 1995
[57]. The MCM-22-based catalyst is very stable.
Cycle lengths in excess of 3 years have been achieved
commercially. The MCM-22 zeolite catalyst is more
monoalkylate selective than large pore zeolites including zeolites beta and Y. This allows the process
to use low feed ratios of benzene to ethylene. Typical
benzene to ethylene ratios are in the range of 3 to
5. The lower benzene to ethylene ratios reduces the
benzene circulation rate which, in turn, improves the
efficiency and reduces the throughput of the benzene
recovery column. Because the process operates with
a reduced benzene circulation rate, plant capacity
can be improved without adding distillation capacity.
This is an important consideration, since distillation
column capacity is a bottleneck in most ethylbenzene
process units. The EBMax process operates at low
temperatures, and therefore the level of xylenes in the
ethylbenzene product is very low, typically less than
10 ppm.
In the EBMax process, benzene is fed to the bottom of the liquid-filled multi-bed reactor. Ethylene is
co-fed with the benzene and also between the catalyst

beds. Polyethylbenzenes (PEB), which are almost


exclusively diethylbenzenes, undergo transalkylation
with benzene in a second reactor. MobilBadger
offers both liquid phase and vapor phase transalkylation processes. The vapor phase process removes
benzene feed co-boilers such as cyclohexane and
methylcyclopentane as well as propyl and butylbenzenes. Because the EBMax process produces very
low levels of propyl and butylbenzenes, for most
applications, the more energy efficient liquid phase
process is preferred. Worldwide, there are currently
10 licensed EBMax units with a cumulative ethylbenzene production capacity of 5 million metric tons
per year.
1.4. LummusUOP process
In 1989, UnocalABB Lummus Crest introduced
a liquid-phase process based on a modified zeolite
Y catalyst. The process design is very similar to the
MobilBadger EBMax process, and uses two reactors, one for benzene alkylation and the other for
diethylbenzene transalkylation. The reactors operate
at close to the critical temperatures of the reaction
mixtures in order to maximize the yield. Reaction
temperatures are typically less than 270 C. Operating pressure is approximately 3.8 MPa (37 bar).
UOP acquired the right to the Unocal patents in
this area when it purchased the technology rights in
1997. ABB Lummus Global and UOP now license
the process. This liquid phase process has several
licensed units with a total ethylbenzene capacity of
2.8 million metric tons per year. It is believed that
the LummusUOP liquid phase ethylbenzene technology has recently shifted to the use of catalysts
based on zeolite beta. In 1994, Lummus also acquired
worldwide licensing rights to a dilute ethylene-based
process developed in China by SINOPEC and the
Dalian Institute of Chemical Physics. The process,
which is based on a ZSM-5 type zeolite, is currently operating in a 30,000 metric ton per year unit in
China [8].
1.5. CDTech process
The CDTech process uses the companys proprietary catalytic distillation (CD) technology [9].
CDTech is a partnership between ABB Lummus

T.F. Degnan Jr. et al. / Applied Catalysis A: General 221 (2001) 283294

Global and Chemical Research & Licensing. Catalyst


Recovery Inc. (CRI), an affiliate of Shell, acquired
the right to CDTechs technology when it purchased
the company in 1997. CDTechs CD column consists
of two sections. The upper section is packed with the
required catalysts and the lower section is equipped
with distillation trays. Ethylene is fed into the CD
column at the bottom of the catalyst bed and benzene
is introduced to the top of the column via the reflux
drum. The counter-current flow reduces coke formation and limits the production of unwanted byproducts. The catalyst is packed in specially designed
bales of steel mesh and fiberglass fabric. These bales
are placed in a column similar to structured packings.
The CDTech process uses zeolite Y as a catalyst [10].
Fig. 4 is a schematic description of the reactive distillation unit. Operating temperatures are reportedly
lower than the LummusUOP process. The product
removed from the bottom of the column is a mixture
of ethylbenzene and PEB. Ethylbenzene is removed
in a separate column and the PEB is sent to a separate
transalkylation reactor. The principal attributes of the
CDTech process reportedly are its long catalyst life
and high product selectivity. There are two commercial units in operation with a total capacity of more
than 850,000 metric tons per year.

287

1.6. Albene process


In 1992, the Indian Petrochemicals Corporation
(IPCL) introduced a process that uses dilute ethanol
rather than ethylene as the alkylating agent. Ethanol
has no distinct advantage over ethylene except that it
can be derived from waste streams or from biomass.
The process reportedly uses [Fe]ZSM-5 as the catalyst. There are no reported commercial units that use
this process.
1.7. Dow Chemicals dilute ethylene process
Dow Chemical has recently introduced a dual
stage ethane dehydrogenation to benzene alkylation
ethylbenzene process [11]. This process mimics the
dilute ethylene process referred to earlier. Ethane is
dehydrogenated over a mordenite catalyst that contains zinc or gallium or one of the platinum group
metals. In one Dow Chemical patent example [12],
gallium-exchanged mordenite produced 86% ethylene selectivity at 50% conversion per pass. The same
patent cites a number of zeolites as possible alkylation catalysts including beta, Y and ZSM-5. The Dow
Chemical process allows for the use of low cost dilute
ethylene without the need to be directly tied to an
FCC unit. Compared to the conventional FCC derived
dilute ethylene processes, this process eliminates concerns about dienes and acetylene, which are known
alkylation catalyst poisons. Because pretreatment provisions are reduced or eliminated, this process can be
more cost effective. There are no published reports of
commercial application of this process.
1.8. Other processes
In 1996, DSM introduced a process for converting
butadiene into ethylbenzene. Chiyoda acquired rights
to the process for license to third parties [13]. The
two-step process converts butadiene into vinylcyclohexene and then into ethylbenzene. No details regarding the catalyst have been made public. It is not known
whether the process is practiced commercially.
1.9. Cumene processes

Fig. 4. Schematic representation of the reactor/distillation column


used in the CDTech ethylbenzene process (Chen [9]).

The primary alkylation and transalkylation reactions in the synthesis of cumene are shown in Fig. 5.

288

T.F. Degnan Jr. et al. / Applied Catalysis A: General 221 (2001) 283294

Fig. 5. Primary alkylation and transalkylation chemistry for cumene production.

Fig. 6 shows the principal side reactions. While the


mechanisms are quite similar to ethylbenzene synthesis, oligomerization is much more important in
cumene synthesis since propylene reacts 2 to 3 orders
of magnitude faster than ethylene over Bronsted acid
catalysts.

Prior to 1992, virtually all cumene was produced


by propylene alkylation of benzene using either solid
phosphoric acid (SPA) or aluminum chloride as catalysts. The SPA process, currently licensed by UOP,
was developed in the 1940s primarily to produce
cumene for aviation fuels [14,15]. More than 40 SPA

Fig. 6. Side reactions typical of the acid catalyzed reaction of benzene with propylene.

T.F. Degnan Jr. et al. / Applied Catalysis A: General 221 (2001) 283294

289

Fig. 7. Flow schematic representation of the SPA process.

plants have been licensed worldwide. The SPA catalyst consists of a complex mixture of orthosiliconphosphate, pyrosiliconphosphate, and polyphosphoric
acid supported on kieselguhr. To maintain the desired
level of activity, small amounts of water are continuously fed into the reactor. The water continually liberates H3 PO4 causing some downstream corrosion. SPA
process conditions include pressures that range from
3.0 to 4.1 MPa, (2940 bar), temperatures that range
from 180 to 230 C, benzene:propylene ratio from 5
to 7, and WHSVs that are between 1 and 2. Approximately 45 wt.% of the product consists of di- and
tri-isopropylbenzenes. A schematic description of the
SPA process is shown in Fig. 7.
In the early 1980s, Monsanto introduced an AlCl3
process based on the same chemistry used in the
ethylbenzene process. This process can be operated at
lower benzene:propylene ratio than the SPA process
because AlCl3 can transalkylate the polyalkylated
benzenes back to cumene. The process also operates
at temperatures lower than the SPA process because
the more highly acidic anhydrous AlCl3 tends to produce significantly more undesired n-propylbenzene at
equivalent temperatures. This technology is currently
used in five plants.

Over the last 7 years, cumene producers have begun to convert to the more environmentally friendly
and more efficient zeolite-based processes. Principal
among these are processes licensed by Dow, CDTech,
MobilBadger, Enichem and UOP. The zeolite-based
processes produce higher cumene yields than the conventional SPA process because most of the diisopropylbenzene (DIPB) byproduct is converted to cumene
in separate transalkylation processes. Operating and
maintenance costs are reduced because there is no
corrosion associated with the zeolite catalysts. Finally,
environmental concerns associated with the disposal
of substantial amounts of phosphoric acid laden kieselguhr or AlCl3 are eliminated by the use of zeolites because they are regenerable and can be safely disposed
of through digestion or landfilling. Table 2 summarizes the manufacturing volume of each of the licensed
zeolite process technologies for cumene manufacture.
1.10. MobilBadger cumene process
First commercialized at Georgia Gulf, Pasadena, TX
plant in 1994, the MobilBadger Cumene process consists of a fixed-bed alkylator, a fixed-bed transalkylator
and a separation section [16,17]. Fig. 8 is a schematic

290

T.F. Degnan Jr. et al. / Applied Catalysis A: General 221 (2001) 283294

Table 2
Commercial zeolite-based cumene units
Process

Year first announced

No. of commercial units

Licensed capacity (million metric ton per year)

MobilBadger
UOP Q-Max
CDTech CDCumene
Enichem
Dow Chemical 3-DDM

1993
1995
1995
1995
1992

7
4
2
1
1a

3125
625
440
265
410

Zeolite catalyst is used in the transalkylation unit only.

description of the process. Fresh and recycled benzene


are combined with liquid propylene in the alkylation
reactor where the propylene is completely reacted. Recycled polyisopropylbenzenes (PIPB) are mixed with
benzene and sent to the transalkylation unit to produce
additional cumene. Trace impurities are removed in
the depropanizer column. Byproduct streams consist
of LPG (mainly propane contained in the propylene
feedstock) and a small residue stream, which can be
used as fuel oil.

As Table 3 shows, the MobilBadger process produces very pure cumene, 99.97 wt.% at 99.7 wt.%
yield. Ethylbenzene, propylbenzene, and butylbenzene levels are an order of magnitude lower than
those obtained with the SPA catalyzed process. The
high cumene purity is primarily attributable to the
high monoalkylation selectivity of the MCM-22 catalyst. This catalyst is unique in its ability to minimize
propylene oligomerization while still exhibiting very
high activity for benzene alkylation. Commercially,

Fig. 8. The MobilBadger liquid phase cumene process flow schematic.

T.F. Degnan Jr. et al. / Applied Catalysis A: General 221 (2001) 283294
Table 3
Product properties from MobilRaytheon cumene process [17,18]
Other zeolites
Cumene (wt%)
Ethylbenzene (ppm)
n-Propylbenzene (ppm)
Butylbenzene (ppm)
Bromine index (ppm)

99.90
200
300
20
50

MCM-22
99.9599.97
25
200
15
<5

the catalyst has demonstrated cycle lengths in excess


of 2 years. Ultimate catalyst life is in excess of 5 years.
The MobilBadger process is used in seven commercial units worldwide. The seven plants have a combined cumene capacity of 3.1 million metric tons per
year, or approximately 50% of the worldwide cumene
capacity. An additional three plants are scheduled to
start up with this process by 2001. This will increase
the total licensed capacity to 4.1 million metric tons.
The selectivity of the MCM-22 catalyst reduces the
size requirements for the fractionation section, and
also reduces the coke forming reactions, which tend to
shorten the cycle length. A key result of the catalysts
low oligomerization selectivity is the ability to design
for low benzene-to-propylene ratiosin the range of
24M. Table 4 compares the selectivities and activities of MCM-22 and zeolite beta zeolites in single
pass alkylation of benzene with propylene. The zeolite
beta catalyst produces significantly more propylene
oligomer, but has a slightly lower selectivity for PIPB.

Table 4
Comparison of MCM-22 and zeolite beta catalysts for cumene
synthesis
MCM-22

Zeolite beta

Propylene
Benzene:propylene (mole ratio)

98.0
3

94.4
3

Total product selectivity (wt%)


Propylene oligomers,
Cumene
PIPB

1.7
84.9
13.4

9.2
79.1
11.7

PIPB/cumene (M, %)

11.3

10.8

Propylene balance, liquid product (wt.%)


Propylene oligomers
4.7
Alkylated aromatic products
95.3

21.4
78.6

conversiona

a Propylene conversion is not complete since this is a once-thru


test at high liquid hourly space velocity (LHSV).

291

1.11. Enichem process


The Enichem cumene process was announced in
1995 and then commercialized at Enichems Porto
Torres, Sardinia plant in March 1996 [19]. The process operates at a benzene to propylene ratio of 4:1.
Cumene yields and selectivities are both greater than
99%. Enichems process uses a modified beta catalyst [10,20]. The process is reported to be readily
retrofit into existing cumene units. Few additional details about this process have been released.
1.12. UOP Q-Max process
UOPs Q-Max process was first commercialized at
BTL Specialty Resin, Blue Island, IL plant in 1996
[21]. The process has also been licensed to Chevron
where it has been in use at Chevron, Port Arthur, TX
plant since 1997 [22,23]. The flow scheme is virtually identical to the MobilRaytheon process design
shown in Fig. 8. UOP has not revealed the composition
of the catalyst used in its process. In fact, very little
additional information is available about this process.
UOP is reportedly working on a MgAPSO-31 catalyst that is very active and selective for cumene [10].
By adjusting the level of Mg in the framework, UOP
has been able to optimize the acidity and reportedly
eliminate the formation of n-propylbenzene.
1.13. Dow chemical 3-DDM process
Dows 3-DDM cumene process uses a highly dealuminated mordenite catalyst in a two-stage fixed-bed
alkylation and transalkylation process [2426]. The
mordenite-based transalkylation stage of this process has operated commercially since 1992 at Dows
Terneuzen, The Netherlands plant [27]. Alkylation
is carried out in a multi-bed fixed-bed reactor. Heat
from the reaction is recovered through feed-effluent
heat exchangers.
Dow has been able to circumvent the coking problems typically associated with mordenite through its
proprietary dealumination process. Dealumination
produces a system of mesopores that change the effective unidimensional pore system of mordenite into
a pseudo-three-dimensional system. The generation
of other means to access the 12-membered ring pores
retards the ability of coke to block access to the active

292

T.F. Degnan Jr. et al. / Applied Catalysis A: General 221 (2001) 283294

sites. Dows mordenite component reportedly has a


silicon-to-aluminum ratio of 70 corresponding to a
zeolite with a low acid site density, but high acid site
strength [10]. The low acid site density reduces the
rates of hydrogen transfer and propylene oligomerization, both of which typically accelerate coke formation. The high acid site strength allows the reaction to
be run at temperatures below which n-propylbenzene
formation becomes significant. The first stage, which
produces cumene and significant levels of DIBP, operates at temperatures below 170 C. Because of the
shape selectivity of the first stage mordenite catalyst,
p-DIPB is the principal C12 isomer in the product.
Triisopropylbenzene makes are very low.
In the second stage, PIPB are transalkylated with
benzene at a reactor temperature of approximately
150 C. DIPB conversions of up to 65% with greater
than 90% selectivity toward cumene have been reported. The process operates at lower benzene-to-propylene ratios than the conventional SPA process.
1.14. CDTech CDCumene process
CDTechs CDCumene process involves the application of CD in a process that uses virtually the same
design as the aforementioned CDTech ethylbenzene
process [9,28,29]. CD has the effect of raising the
benzene-to-propylene ratio in the bulk phase surrounding the catalyst. Cumene yields in excess of
99% are claimed for this dual stage process. Catalyst
lifetimes are projected to be greater than 2 years in
commercial operations based on pilot plant tests that
have lasted over 6000 h on a single catalyst charge.
Alkylation is carried out isothermally at relatively
low temperatures and pressures of approximately
5 bar in the overhead. The mild reaction conditions
are reported to produce cumene of very high purity. The main impurities are reportedly ethylbenzene
(220 ppm) and n-propylbenzene (350 ppm). As in
Mobils and Dows processes, DIPB produced in the
alkylation stage are converted to cumene in a second
stage transalkylation reactor.
The nature of CDTechs alklyation and transalkylation catalysts has not been disclosed, although
CDTech patents refer to zeolites Y, omega and beta.
Till date, there is one operating unit which is licensed
to GPOrgetelko-Dzeryinsk, in Nizhny Novgrod,
Russia [9]. The unit has a total cumene capacity

of 170 million metric tons per year. Another unit in


Taiwan with a capacity of 240 million metric tons, is
slated to be commissioned in 2000.
A related technology, CDTechs Catstill process,
combines CDTechs cumene and ethylbenzene processes with the objective of removing benzene from
reformate by alkylation with FCC off-gas.
1.15. Other processes
LummusUnocal reportedly have been developing
a cumene process that is similar to their commercial
ethylbenzene process. The process, which has been run
on a pilot-plant scale, uses a modified trilobe Y zeolite
catalyst that produces improved product selectivities
[10].
In other laboratory scale cumene process studies, Du Ponts nafionsilica composites have shown
very good activity for liquid phase alkylation under very mild conditions. The catalyst consists of
nanometer-sized nafion particles dispersed in a highly
porous silica matrix [30].
1.16. Future developmentsethylbenzene
The approximately 40% of the worlds ethylbenzene capacity that still uses AlCl3 is testimony to the
efficiency and economy of this process. To capture this
segment of the industry, zeolite catalysts will have to
be developed that operate at close to the same very
low benzene-to-ethylene ratios that make the AlCl3
process economically attractive. Heat management in
fixed-bed reactors becomes a design concern at the low
benzene-to-ethylene ratios that characterize the AlCl3
process. Hence, process and catalyst innovations will
have to evolve concurrently to achieve the goal of low
benzene-to-ethylene ratios.
Another opportunity for advancement in ethylbenzene synthesis is in the development of liquid phase
processes that can handle low cost feedstocks including dilute ethylene such as ethaneethylene mixtures.
The use of dilute ethylene has become increasingly
attractive since it has the potential to debottleneck
ethylene crackers. Currently, higher temperature, vapor phase technologies can tolerate contaminants that
enter with the dilute ethylene feed from FCC units.
However, these same contaminants can accelerate
catalyst aging in lower temperature, liquid phase

T.F. Degnan Jr. et al. / Applied Catalysis A: General 221 (2001) 283294

operations because they are more strongly adsorbed


at the lower temperatures. Acid catalysts that tolerate elevated levels of contaminants would facilitate
the development of dilute ethylene-based processes.
These same catalysts could be useful in applications
where lower cost or lower quality benzene feeds are
all that are available.
Future ethylbenzene alkylation catalyst development efforts will also undoubtedly focus on systems
that convert less conventional feedstocks including
ethane and ethanol. The two-step Dow ethane-based
process for ethylbenzene production is believed to be
uneconomical because of its high capital investment
requirement [31]. However, it is a very attractive concept, and could be implemented if more efficient catalysts or improved process designs could be developed.
There are similar opportunities for improved
transalkylation catalysts and processes. Heavies made
with the current zeolite-based transalkylation catalysts are already very low, however, improved catalysts could lead to even lower byproduct makes and
yield losses due to residue formation. The majority of
the reaction by-products in the zeolite-catalyzed processes are produced in the transalkylator. There are
opportunities to develop more active transalkylation
catalysts that allow operation at lower temperatures
where by-product formation can be minimized.
Process development efforts will likely focus on increased integration of the alkylation and transalkylation stages. The two processes frequently operate at
different temperatures. The challenge here is to match
activities and aging rates. Other efforts will likely focus on improved process heat integration.
1.17. Future developmentscumene
It is virtually inevitable that zeolite catalyzed, liquid
phase production of cumene will eventually displace
the SPA and AlCl3 based processes. The situation is
not analogous to ethylbenzene, since the predominant
competing technology that zeolite processes face is the
far less efficient and environmentally undesirable than
SPA process. The current zeolite-based processes allow for significant debottlenecking of SPA plants with
little additional capital. More importantly, the zeolite
processes have been able to produce cumene with purities that cannot be approached by either the SPA
or the AlCl3 process. Tighter product specifications,

293

brought about by the more selective zeolite processes


will drive all manufacturers to select one of the aforementioned zeolite technologies, possibly in the very
near term.
Additional research will undoubtedly be directed towards identifying improved alkylation and transalkylation catalysts. Alkylation catalysts that operate at
benzene-to-propylene ratios of 3 or possibly even
lower are desirable because they greatly reduce the
need for costly separation capacity and because they
can debottleneck existing units. Transalkylation and
alkylation catalysts that produce even lower byproduct
concentrations at low benzene to propylene levels are
likely to be the focus of future research and development efforts. The challenge of combining both lower
by-product make and lower benzene-to-propylene
feed ratios involves overcoming higher aging rates
implicit in using higher propylene feed concentrations and identifying more monoalkylation selective
catalysts with activity similar to todays zeolite catalysts. Materials that surpass even MCM-22s superior
ability to operate at high propylene concentrations
without catalyzing excessive oligomerization could
lead to important advances in cumene synthesis.
Current zeolite catalysts already operate at process
temperatures that require minimal external heat addition. Heat integration and heat management will be of
increasing concern at the lower benzene-to-propylene
ratios because the cumene synthesis reaction is so
highly exothermic (H f = 98 kJ mol1 ). Recycle,
particularly in the alkylation reactor is likely to become increasingly important as a heat management
strategy. The key will be how to limit the build-up
of by-products and feed impurities in these recycle
loops, particularly as manufacturers seek cheaper and
consequently lower quality feedstocks. As in the case
of ethylbenzene, process and catalyst innovations will
have to develop concurrently.
Recently, there have been several process developments that allow for the production of phenol
without co-production of acetone. These would tend
to compete directly with the need to produce cumene.
Developments include the AlphOx process, a joint
development of Solutia and the Boreskov Institute of
Catalysis [3235] which allows for the direct oxidation of benzene to produce phenol. Economic analyses
have shown that these are attractive only in specific instances where, e.g. a cheap source of N2 O is available.

294

T.F. Degnan Jr. et al. / Applied Catalysis A: General 221 (2001) 283294

Nevertheless, these developments have shown that


direct oxidation is possible and further innovations in
this area should probably be expected. The demands
for acetone and phenol have generally trended to follow each other. However, as bisphenol A becomes an
even more important end use for phenol and acetone,
there will be a need for a separate source of phenol.
The synthesis of bisphenol A requires 2 mol of phenol
for every 1 mol of acetone, while the peroxidation of
cumene produces 1 mol of each. Still, processes such
as the direct oxidation of benzene are unlikely to have
a major impact on cumene demand in the short term
since there are competing processes such as Mitsuis
for converting acetone back to propylene.
There is also the prospect for increased demand
for some of the cumene by-products such as DIPB.
The production of diphenols from DIPB is important
for synthesis of resorcinol (from m-DIPB) and hydroquinone (from p-DIPB). It is likely that the market may
soon see the introduction of cumene-based processes
that specifically produce these isomers with high degrees of selectivity.
References
[1] M.A. Caesar, SRI Consulting PEP, Report No. 219, 1999 p. 2.
[2] F.G. Dwyer, in: W.R. Moser (Ed.), Catalysis of Organic
Reactions: Chemical Industries, Marcel Dekker, New York,
1981, p. 39.
[3] N.Y. Chen, W.A. Garwood, F.G. Dwyer, Shape Selective
Catalysis in Industrial Applications, Marcel Dekker, New
York, 1996, p. 208.
[4] F.G. Dwyer, S. Ram, AIChE Spring National Meeting,
Houston, TX, 711 April 1991.
[5] B.R. Maerz, C.R. Venkat, S.S. Chen, D.N. Mazzone, DeWitt,
1996 Petrochemical Review, Houston, TX, 321 March 1996.
[6] D.N. Mazzone, P.J. Lewis, C.R. Venkat, B.R. Maerz,
Hydrocarbon Asia 7 (3) (1997) 56.
[7] J.C. Cheng, T.F. Degnan, J.S. Beck, Y.Y. Huang, M.
Kalyanaraman, J.A. Kowalski, C.A. Loehr, D.N. Mazzone,
Stud. Surf. Sci. Catal. 105 (1998) 56.
[8] Chemical Market Report, 17 October 1994, p. 27.

[9] J. Chen, in: Proceedings of the Worldwide Solid Acid Process


Conference 1993, Houston, TX, November 1993, Catalyst
Consultants Inc., Spring House, PA, 1993.
[10] J.A. Horsley, Chemtech 10 (1997) 45.
[11] European Chemical News, 69, 18 May 1998, p. 18.
[12] R.F. Progue, World Patent 96/34843, Dow Chemical
Company, 1996.
[13] Chemical Week, 16 October 1996, p. 28.
[14] V.N. Ipatieff, US Patent 2,382,318, 1945.
[15] P.R. Pujado, J.R. Salazar, C.V. Berger, Hydrocarbon Process.
55 (3) (1976) 91.
[16] Chemical Week, 14 June 1995, p. 16.
[17] Hydrocarbon Processing, October 1996, p. 40.
[18] Data from Georgia Gulf cumene unit evaluation, June 1996.
[19] C. Perego, Chim. Ind. (Milan) 80 (1998) 584.
[20] G.R. Meima, M.J.M. van der Aalst, M.S.U. Samson, J.M.
Garces, J.G. Lee, Edrol Erdgas Kohle 7/8 (1996) 315.
[21] M.F. Bentham, G.J. Gadja, R.H. Jensen, H.A. Zinner, Erdoel
Erdgas Kohle 113 (1997) 84.
[22] Oil Gas J. 95 (2) 29; 8 September (1997) 91.
[23] Chemical Week, 8 January (1997) 17.
[24] G.R. Meima, M.J.M. van der Aalst, M.S.U. Samson, N.J.
Vossenberg, R.J.O. Adriaansens, J.M. Garces, G.J. Lee, in:
Proceedings of the Worldwide Solid Acid Process Conference
1993, Catalyst Consultants, Inc., Spring House, PA, 1993.
[25] J.E. Wallace, K.V. Shah, DeWitt Petrochemical Review, 2224
March 1994, Houston, TX, pp. D16.
[26] G.R. Meima, M.J.M. van der Aalst, M.S.U. Samson, J.M.
Garces, J.G. Lee, in: R. von Ballmoos, J.B. Higgins,
M.M.J. Treacy (Eds.), Proceedings of the 9th International
Zeolite Conference, Vol. 2, Montreal, Canada, 1992;
Butterworth-Heinemann, Boston, 1993, p. 327.
[27] G.R. Meima, CATTECH 2 (1) (1998) 5.
[28] A. Sy, J. Smith, J. Chen, F. Dautzenberg, DeWitt
Petrochemical Review, Houston, TX, 2325 March 1993,
pp. Q128.
[29] J.A.O. Portela, J.M. Hildreth, DeWitt Petrochemical Review,
Houston, TX, 2224 March 1994, Paper E122.
[30] M.A. Harmer, W.E. Farneth, Q. Sun, J. Am. Chem. Soc. 118
(1996) 7708.
[31] S. Naqvi, Process Economics Program (PEP), Review 97-3,
SRI Consulting, Menlo Park, CA, 1998.
[32] G.I. Panov, Appl. Catal. A: Gen. 98 (1993) 1.
[33] A.K. Uiarte, 3rd World Congress on Oxidation Catalysis, San
Diego, CA, September 1997.
[34] K. Walsh, Chem. Week 18 (1998) 10.
[35] Chemical Market Report, 30 December 1996, p. 250.

You might also like