You are on page 1of 14

Ethylbenzene

Ethylbenzene
Vincent A. Welch, Washington Group Intl., Inc. (formerly The Badger Company), Cambridge, Massachusetts,
United States (Chap. 1, 2, 3, 4, 5, 6, 7, 8 and 9)
Kevin J. Fallon, Washington Group Intl., Inc. (formerly The Badger Company), Cambridge, Massachusetts,
United States (Chap. 1, 2, 3, 4, 5, 6, 7, 8 and 9)
Heinz-Peter Gelbke, BASF Aktiengesellschaft, Ludwigshafen, Federal Republic of Germany (Chap. 10)

Introduction . . . . . . . . . . . . . . . . .
Physical Properties . . . . . . . . . . . .
Chemical Properties . . . . . . . . . . .
Production . . . . . . . . . . . . . . . . . .
Alkylation with Non-Zeolite Lewis
Acid Catalysts . . . . . . . . . . . . . . .
4.2. Vapor-Phase Alkylation over Zeolites
4.3. Liquid-Phase Alkylation over Zeolites
4.4. Mixed-Phase Zeolite-Based Process .

1.
2.
3.
4.
4.1.

1
2
2
2

4.5.
5.
6.
7.

3
5
7
8

8.
9.
10.
11.

1. Introduction
Ethylbenzene
[100-41-4],
phenylethane,
C6 H5 CH2 CH3 , M r 106.168, is a single-ring
alkylaromatic compound. It is almost exclusively (>99 %) used as an intermediate for
the manufacture of styrene monomer [100-425], C6 H5 CH=CH2 , one of the most important
large-volume commodity chemicals. Styrene
production, which uses ethylbenzene as a starting raw material, consumes ca. 50 % of the
worlds benzene production. Less than 1 % of
the ethylbenzene produced is used as a paint
solvent or as an intermediate for the production
of diethylbenzene and acetophenone.
Commercially, almost all ethylbenzene is
produced by alkylating benzene with ethylene.
The newest technologies utilize synthetic zeolites installed in xed-bed reactors to catalyze the alkylation in the liquid phase. Another
proven route uses narrower pore synthetic zeolites, also installed in xed-bed reactors, to effect
the alkylation in the vapor phase. A considerable quantity of ethylbenzene is still produced by
alkylation with homogeneous aluminum chloride catalyst in the liquid phase, though the recent trend in the industry has been to retrot
such units with zeolite technology. The alkylation of aromatic hydrocarbons with olens in
c 2005 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

10.1002/14356007.a10 035.pub2

Separation from Mixed C8 Streams


Environmental Protection . . . . . . .
Quality Specications . . . . . . . . . .
Handling, Storage, and
Transportation . . . . . . . . . . . . . .
Uses . . . . . . . . . . . . . . . . . . . . .
Economic Aspects . . . . . . . . . . . .
Toxicology . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . .

.
.
.

8
9
9

.
.
.
.
.

10
10
10
11
13

the presence of aluminum chloride catalyst was


rst practiced by Balsohn in 1879 [1]. However,
Friedel and Crafts pioneered much of the early
research on alkylation and aluminum chloride
catalysis.
Several facilities built during the 1960s recovered ethylbenzene by fractionation of mixed
xylenes produced in renery catalytic reforming
units. This practice has largely been discontinued due to poor economics that result from high
energy and investment costs, as well as small
economies of scale in comparison to the conventional alkylation routes.
Ethylbenzene was rst produced on a commercial scale in the 1930s by Dow Chemical
in the United States and by BASF in Germany. The ethylbenzene/styrene industry remained relatively insignicant until World
War II. The tremendous demand for synthetic
styrene butadiene rubber (SBR) during the war
prompted accelerated technology improvements
and tremendous capacity expansion. This considerable wartime effort led to the construction
of several large-scale factories, and styrene production quickly became a signicant industry. In
1999 world annual production capacity of ethylbenzene reached almost 25 106 t. Throughout the 1990s, most of the capacity increase oc-

Ethylbenzene

curred in the Far Eastern countries other than


Japan, where the basic petrochemical industries
have undergone considerable development and
expansion.

2. Physical Properties
Under normal conditions, ethylbenzene is a
clear, colorless liquid with a characteristic aromatic odor. Ethylbenzene is an irritant to the skin
and eyes and is moderately toxic by ingestion,
inhalation, and skin adsorption. Some physical
properties of ethylbenzene follow [2, 3]:
Density

mp
bp
Refractive index

at 15 C
at 20 C
at 25 C
at 101.3 kPa
at 20 C
at 25 C

Critical pressure
Critical temperature
Flash point (closed cup)
Autoignition temperature
Flammability limit lower
upper
Latent heat
fusion
vaporization
Heating value,
gross
net
Kinematic viscosity at 37.8 C
at 98.9 C
Surface tension
Specic heat capacity
ideal gas, 25 C
liquid, 25 C
Acentric factor
Critical
compressibility

0.87139 g/cm3
0.8670 g/cm3
0.86262 g/cm3
94.949 C
136.2 C
1.49588
1.49320
3609 kPa (36.09 bar)
344.02 C
15 C
460 C
1.0 %
6.7 %
86.3 J/g
335 J/g
42 999 J/g
40 928 J/g
0.6428106 m2 /S
0.390106 m2 /S
28.48 mN/m
1169 J kg1 K1
1752 J kg1 K1
0.3026
0.263

3. Chemical Properties
The most important commercial reaction of
ethylbenzene is its dehydrogenation to styrene.
The reaction is carried out at high temperature (600 660 C), usually over a potassiumpromoted iron oxide catalyst. Steam is used as a
diluent. Commercially, selectivities for styrene
range from 90 to 97 mol % with per-pass conversions of 60 70 %. Side reactions involve
mainly the dealkylation of ethylbenzene to benzene and toluene.

A reaction of increasing commercial importance is the oxidation of ethylbenzene by air to


the hydroperoxide C6 H5 CH(OOH)CH3 [307132-7]. The reaction takes place in the liquid
phase without a catalyst. However, because hydroperoxides are unstable, exposure to high temperature must be minimized to reduce the rate of
decomposition. The production of byproducts is
reduced if the temperature is gradually lowered
during the course of the reaction. The hydroperoxide is subsequently treated with propylene to
give styrene and propylene oxide as co-products.
In 1999 about 15 % of the ethylbenzene produced worldwide was used in the co-production
of styrene monomer and propylene oxide.
Like toluene, ethylbenzene can be dealkylated catalytically or thermally to benzene.
Ethylbenzene also undergoes other reactions
typical of alkylaromatic compounds [4].

4. Production
Alkylation of benzene with ethylene is the
source of nearly all ethylbenzene. For several
decades most alkylation plants utilized dissolved
Lewis acids, in most cases aluminum chloride,
to catalyze the reaction in the liquid phase.
About 40 % of worldwide ethylbenzene production still utilizes variations on this method. Although the aluminum chloride route generally
provides competitive economics, disposal of the
resulting waste streams has become increasingly
costly. In addition, this route is susceptible to severe corrosion of equipment and piping. Since
the early 1980s, technologies using heterogeneous zeolites, operating in the vapor phase and
more recently in the liquid phase, have been
most often selected for new grass-roots plants.
More recently, increasing environmental pressures and improvements in the zeolite processes
have provided incentives to a number of operators of aluminum chloride units to retrot with
zeolite technology. About 106 t of capacity in
several plants were changed to zeolite technology from 1997 to 1999, and as of 2000, more
conversions are in the engineering phase. It has
been nearly ten years since the last signicant
grass-roots capacity utilizing aluminum chloride
technology came on-stream.
A very minor amount of ethylbenzene is produced by superfractionation of mixed C8 aro-

Ethylbenzene

Figure 1. Aluminum chloride process for ethylbenzene production


a) Catalyst mix tank; b) Alkylation reactor; c) Settling tank; d) Acid separator; e) Caustic separator; f) Water separator;
g) Benzene recovery column; h) Benzene dehydrator column; i) Ethylbenzene recovery column; j) Polyethylbenzene column

matic streams. Only a small number of plants in


the world still utilize this method.

4.1. Alkylation with Non-Zeolite Lewis


Acid Catalysts
Liquid-phase aluminum chloride processes were
rst commercialized in the 1930s. Several companies developed variations of this technology,
including Dow Chemical, BASF, Shell Chemical, Monsanto/Lummus, Societe Chimiques des
Charbonnages, and Union Carbide/Badger. Of
the aluminum chloride plants still in operation, the majority of them utilize the Monsanto/Lummus technology, widely considered
the most advanced of the aluminum chloride
processes. Lummus discontinued licensing this
technology with the commercialization of their
liquid-phase zeolite process around 1990.
Alkylation of benzene with ethylene is highly
exothermic (H = 114 kJ/mol); in the presence of aluminum chloride the reaction is very
fast and produces almost stoichiometric yields of
ethylbenzene. In addition to AlCl3 , a wide range
of Lewis acid catalysts, including AlBr3 , FeCl3 ,
ZrCl4 , and BF3 , have been used. Aluminum
chloride processes generally use ethyl chloride
or hydrogen chloride as a catalyst promoter.
These halide promoters reduce the amount of
AlCl3 required. The reaction mechanism has
been studied in detail [5].

In the older conventional AlCl 3 process (see


Fig. 1) which is seldom practiced now, three
phases are present in the reactor: aromatic liquid, ethylene gas, and a liquid catalyst complex phase (a reddish brown material called red
oil). A mixture of catalyst complex, dry benzene, and recycled polyalkylbenzenes is continuously fed to the reactor and agitated to disperse the catalyst complex phase in the aromatic
phase. Ethylene and the catalyst promoter are
injected into the reaction mixture through spargers, and essentially 100 % of the ethylene is converted. Commercial plants typically operate at
ethylene/benzene molar ratios of ca. 0.3 0.35.
As this ratio is increased, more side reactions,
such as transalkylation and isomeric rearrangement, occurr. Further alkylation of ethylbenzene
leads to the reversible formation of lower molecular mass polyalkylbenzenes. The loss in yield
due to residue formation is minimized by recycling this material to the alkylation reactor.
In addition, because the reaction occurs close to
thermodynamic equilibrium, the traditional processes use a single reactor to alkylate benzene
and transalkylate polyalkylbenzenes.
The reaction temperature is generally limited to 130 C; a higher temperature rapidly deactivates the catalyst and favors formation of
nonaromatics and polyalkylbenzenes, which are
preferentially absorbed by the highly acidic catalyst complex and form byproducts. Sufcient
pressure is maintained to keep the reactants in
the liquid phase. Because the reaction mixture

Ethylbenzene

Figure 2. Homogeneous liquid-phase alkylation process for ethylbenzene production


a) Benzene drying column; b) Alkylation reactor; c) Catalyst preparation tank; d) Transalkylator; e) Flash drum; f) Vent gas
scrubbing system; g) Decantor; h) Neutralization system

is highly corrosive, the alkylation reactors are


lined with brick or glass. High-alloy construction materials are also required for the piping
and handling systems.
The liquid reactor efuent is cooled and discharged into a settler, where the heavy catalyst
phase is decanted from the organic liquid-phase
and recycled. The organic phase is washed with
water and caustic to remove dissolved AlCl3 and
promoter. The aqueous phase from these treatment steps is rst neutralized and then recovered
as a saturated aluminum chloride solution and a
wet aluminum hydroxide sludge.
Removal of dissolved catalyst from the organic stream has long been a challenge for ethylbenzene producers. CdF Chimie found that more
complete recovery of AlCl3 could be achieved
by rst contacting the organic phase with ammonia instead of sodium hydroxide [6].
Purication of the ethylbenzene product is
usually accomplished in a series of three distillation columns. The unconverted benzene is
recovered in the rst column as an overhead distillate. The second column separates the ethylbenzene product from the heavier polyalkylated
components. The bottoms product of the second
column is fed to a nal column where the recyclable polyalkylbenzenes are stripped from non-

recyclable high molecular mass residue compounds. The residue stream, or ux oil, consisting primarily of polycyclic aromatics, is burned
as fuel.
Because the alkylation mixture can tolerate
only minor amounts of water, the recycle benzene and fresh benzene must be dried thoroughly
prior to entering the reactor. Water not only increases corrosion, but also decreases catalyst activity. Benzene dehydration is accomplished in
a separate column.
The improved Monsanto process (see Fig.
2) has several advantages over the conventional
AlCl3 processes. Over time many conventional
AlCl3 plants have been retrotted with this technology. The most important advantage is a signicant reduction in the amount of AlCl3 catalyst used, thus lessening the problem and cost
of waste catalyst disposal. Monsanto found that
by increasing temperature and by careful control
of ethylene addition, the required AlCl3 concentration could be reduced to the solubility limit,
thereby eliminating the separate catalyst complex phase [7]. Therefore, the alkylation occurs
in a single homogeneous liquid phase instead
of in the two liquid phases of earlier processes.
Monsanto claimed that a separate catalyst complex phase may actually prevent the attainment

Ethylbenzene
of maximum reactor yields. With a few exceptions, the ow scheme of the Monsanto process
is nearly the same as that of more traditional processes. The process is also capable of operating
with low-concentration ethylene feed. Typically,
the alkylation temperature is maintained at 160
180 C. This higher operating temperature enhances catalyst activity with the additional benet that the heat of reaction can be recovered as
low-pressure steam.
Whereas the traditional process accomplishes alkylation and transalkylation in a single reactor, the homogeneous catalyst system
must employ a separate transalkylation reactor.
At the lower catalyst concentrations, the recycle
of substantial amounts of polyalkylbenzenes terminates the alkylation reaction. Therefore, only
dry benzene, ethylene, and catalyst are fed to
the alkylation reactor. The recycle polyethylbenzene stream is mixed with the alkylation reactor
efuent prior to entering the transalkylation reactor. The transalkylation reactor is operated at
much lower temperature than the primary alkylation reactor.
After transalkylation, the reaction products
are washed and neutralized to remove residual
AlCl3 . With the homogeneous process, all of the
catalyst remains in solution. The catalyst-free
organic reaction mixture is then puried by using the sequence described for the conventional
AlCl3 process. As with other AlCl3 processes,
the organic residue is used as fuel and the aluminum chloride waste streams are usually sold
or sent to treatment facilities.
In 1999 about 40 % of the ethylbenzene manufactured worldwide used aluminum chloride
technology. Another technology based on Lewis
acids is the so-called Alkar process. Developed
by UOP and based on boron triuoride catalyst,
this process had modest success in the 1960s,
but fell from favor because of high maintenance
costs resulting from the severe corrosion caused
by small quantities of water. In the developed
countries, only one Alkar plant is still in operation.
However, the process produced a high-purity
ethylbenzene product and could use dilute ethylene feedstock. If the entry of water into the
process was strictly prevented, the corrosion
problems associated with aluminum chloride
processes were avoided. However, even small

amounts of water (< 1 g/kg) hydrolyzed the BF3


catalyst.
The alkylation reaction took place at high
pressure (2.5 3.5 MPa) and low temperature
(100 150 C). Dried benzene, ethylene, and
makeup BF3 catalyst were fed to the reactor. Typically, ethylene/benzene molar ratios between 0.15 and 0.2 were used. The reactor inlet
temperature was controlled by recycling a small
portion of the reactor efuent. Transalkylation
took place in a separate reactor. Dry benzene,
BF3 catalyst, and recycled polyethylbenzenes
were fed to the transalkylation reactor, which
operated at higher temperature (180 230 C)
than the alkylation reactor. The efuent streams
from the two reactors were combined and passed
to a benzene recovery column where benzene
was separated for recycle to the reactors. Boron
triuoride and light hydrocarbons were taken
overhead as a vapor stream from which the BF3
was recovered for recycle. The bottoms from the
benzene recovery column was sent to a product column where ethylbenzene of > 99.9 % purity was taken overhead. A nal column served
to recover polyethylbenzenes for recycle to the
transalkylation reactor.
The Alkar process could operate with an ethylene feed containing as little as 8 10 mol %
ethylene, enabling a variety of renery and cokeoven gas streams to be used. However, purication of these streams was necessary to remove
components that poison the BF3 catalyst, e.g.,
trace amounts of water, sulfur compounds, and
oxygenates.

4.2. Vapor-Phase Alkylation over


Zeolites
The Mobil Badger vapor-phase technology
was developed in the 1970s around Mobils
ZSM-5 synthetic zeolite and has been available in several different designs. The original rst-generation design, commercialized by
American Hoechst in 1980, carried out vaporphase alkylation and transalkylation in a single reactor by recycling polyethylbenzene to the
front end of the process, similar to conventional
aluminum chloride technology. The newest,
so-called third-generation technology performs
transalkylation in a separate lower pressure reaction step. The third-generation technology of-

Ethylbenzene

Figure 3. Third-generation Mobil Badger ethylbenzene process


a)Reactor-feed heater; b) Alkylation reactor; c) Benzene recovery column; d) Ethylbenzene recovery column;
e) Polyethylenebenzene recovery column; f) Secondary reactor; g) Stabilizer

fered signicant benets in yield, purity, and


capital cost, and was widely used in the 1990s
to debottleneck older vapor-phase plants.
The vapor-phase zeolite process is particularly suited to dilute ethylene streams, particularly renery off-gas from uid catalytic cracking (FCC) units. Until technologies using zeolites in the liquid phase were commercialized in
the 1990s, the vapor-phase zeolite process was
the dominant technology of new plants, primarily because it avoided the aqueous waste streams
produced from aluminum chloride plants. Mobil
Badger licensed a total of 31 units since 1980,
and the technology is still licensed for diluteethylene-based plants.
The xed-bed ZSM-5 catalyst promotes the
same overall alkylation chemistry as in the other
processes; however, ethylene molecules are adsorbed onto the Brnsted acid sites within the
catalyst, which activates the ethylene molecule
and allows bonding with benzene molecules to
occur. Hence, the range of higher alkylated aromatic byproducts formed by the Mobil Badger
process is somewhat different than those of the
Friedel Crafts processes.
Carbon steel is the primary material of construction; high-alloy materials and brick linings are not required. A diagram of the thirdgeneration design is shown in Figure 3.
The alkylation reactor typically operates in
the range of 350 450 C and 1 3 MPa. At
this temperature, > 99 % of the net process heat

input and exothermic heat of reaction can be recovered as steam. The reaction section includes a
multibed reactor, a red heater and heat recovery
equipment. The reactor operates with signicant
overall excess of benzene relative to ethylene.
Slow catalyst deactivation occurs as a result
of coke formation and requires periodic regeneration. In situ regeneration takes ca. 36 h and
may be necessary after 18 24 months of operation, depending on the operating conditions.
The catalyst is less sensitive to water, sulfur, and
other poisons than the Lewis acid catalysts and
zeolites operating in the liquid phase.
The reactor efuent passes to the purication section as a vapor. The benzene from this
stream is distilled overhead in the rst distillation column and is subsequently recycled to the
reactors. The ethylbenzene product is taken as
the overhead product from the second column.
The bottoms product from this column is sent to
the last column where the recyclable alkylbenzenes and polyalkylbenzenes are separated from
heavy nonrecyclable residue. The low-viscosity
residue stream, consisting mainly of diphenylmethane and diphenylethane, is burned as fuel.
Recyclable higher alkylbenzenes and
polyalkylbenzenes are sent to the vapor-phase
transalkylator, where they are converted in the
presence of excess benzene over zeolite catalyst.
Because the transalkylator has lower pressure
but higher temperature relative to the alkylator,
higher alkylbenzenes are dealkylated while di-

Ethylbenzene

Figure 4. Lummus/UOP ethylbenzene process [9]


a) Alkylation reactor; b) Transalkylation reactor; c) Benzene column; d) Ethylbenzene column; e) Polyethylbenzene column

ethylbenzene is transalkylated to ethylbenzene.


The ability to dealkylate higher alkylbenzenes,
known to be residue precursors, serves to decrease overall residue production.
The older rst- and second-generation processes were similar, the major difference being
that the recycle polyethylbenzene stream was recycled to the alkylation reactor. For that reason
the process had inferior performance relative to
the third-generation design.
Though most often used with polymer-grade
ethylene, this process is adaptable to dilute ethylene feedstocks. The process has operated on
a mixed C2 stream provided from a debottlenecked distillation train from a naphtha cracker.
Additionally, and perhaps of more interest due
to low feedstock cost, is the adaptability to dilute ethylene steams produced from FCC offgas. Two world-scale Mobil Badger units operate with FCC off-gas, one since 1991 and one
since 1998 [8].

4.3. Liquid-Phase Alkylation over


Zeolites
All-liquid phase processes using zeolite catalysts began to appear in commercial operation
in 1990, the rst plant being operated by Nippon SM in Oita, Japan under license from ABB
Lummus Global and Unocal (UOP later became ABBs partner when they acquired Unocals zeolite business). This process has uti-

lized ultrastable zeolite Y or more recently zeolite beta. Another liquid-phase EB process,
EBMax, became available from Mobil Badger and is based on a Mobil MCM-22 catalyst
and rst operated at Chiba Styrene Monomer
Corp., also in Japan. A total of twelve all-liquidphase zeolite plants were operating at the end of
1999. Although there are differences between
the two available liquid-phase technologies, the
latest versions of both lead to lower investment
cost and better product quality than was possible with previously available technologies using polymer-grade ethylene. The vapor-phase
technology is now typically licensed by Mobil/Badger for dilute-ethylene applications.
These liquid phase processes all utilize wider
pore zeolites than ZSM-5, which is necessary
to overcome the diffusional limitations of the
liquid-phase mechanism. Both licensors usually
recommend off-site regeneration of catalyst because of long catalyst cycle times. Off-site regeneration results in less on-site equipment being required and hence in reduced investment.
The owsheets of the two technologies are
quite similar (Figs. 4 and 5). Ethylene is injected
into a xed bed alkylation reactor with multiple
stages in the presence of excess benzene. Reactor temperatures vary from process to process,
but they must be kept below the critical temperature of benzene (289 C). Pressures must
be high enough to keep the light gases in solution and are on the order of 4 MPa. Excess ben-

Ethylbenzene

Figure 5. Mobil Badger EBMax process


a) Alkylation reactor; b) Transalkylation reactor; c) Benzene column; d) Vent-gas column; e) Ethylbenzene column;
f) Polyethylbenzene column

zene is distilled overhead from the efuent and


is recycled to alkylation. The benzene column
bottoms enter the ethylbenzene product column,
from which the ethylbenzene is distilled overhead. The bottoms of the ethylbenzene column
are fed to the polyethylbenzene column, which
separates higher alkylbenzenes and polyethylbenzenes from the residue. This overhead stream
is fed to the liquid-phase transalkylation reactor,
where it is combined with excess benzene from
the benzene column distillate. Efuent from the
transalkylator is returned to the distillation train.
Besides benzene, major impurities in the
ethylbenzene product can include nonaromatics (naphthenes), toluene, and higher alkylbenzenes. Depending on the operating conditions
and technology, these components can originate
from the feed benzene or be generated in the
reactor. There is a wide variation in distillation
column operating conditions, particularly in revamped facilities. There is also a wide variation
in excess benzene from plant to plant. These details are held condentially by the licensors.

4.4. Mixed-Phase Zeolite-Based Process


A mixed-phase ethylbenzene process is offered
for license by CDTech, a partnership between
ABB Lummus Global and Chemical Research
and Licensing, Co. [10, 11] The rst unit started
in 1994, and as of 1999, three plants were op-

erating. The main distinction of this process is


the alkylation reactor, which contains bales of
zeolite catalyst in a reactive-distillation column.
Ethylene gas and benzene liquid are fed to the
reactive-distillation column. Because of its ability to handle the ethylene feed in the vapor phase,
the process has been applied to dilute ethylene
streams produced from steam cracker distillation trains. It has also been applied to polymer
grade ethylene. A ow scheme is shown in Figure 6. An alkylator and benzene stripper operate
together as a distillation column. The overhead
benzene and unconverted ethylene are in turn fed
to a nishing reactor, which also utilizes zeolite. Bottoms from the benzene stripper are fractionated into ethylbenzene product in the overhead product column, and then transalkylatable
polyethylbenzenes are distilled from the residue.
The polyethylbenzenes are sent to a liquid-phase
transalkylation reactor in the presence of excess
benzene, and the transalkylation efuent is returned to the fractionation train.

4.5. Separation from Mixed C8 Streams


Less than 1 % of worldwide ethylbenzene production is recovered from mixed xylene streams,
usually in conjunction with xylene production
from reformate. Although adsorption processes
have been developed, most notably the EBEX
process of UOP, ethylbenzene production from

Ethylbenzene

Figure 6. CDTech ethylbenzene process


a) Finishing reactor; b) Transalkylator; c) Alkylator; d) Benzene stripper; e) Ethylbenzene column; f) Polyethylbenzene column
BFW = boiler feed water, PEB = polyethylbenzene

these sources has been mainly performed by distillation. Because of the difculty of the separation, the process is generally termed superfractionation. It was rst undertaken by Cosden Oil
& Chemical Company in 1957, using technology jointly developed with the Badger Company. The separation generally requires three
distillation columns in series, each having over
100 stages. Several units were built during the
1960s in the United States, Europe, and Japan.
However, the increased cost of energy and high
capital cost has made this route noncompetitive.

(Class 1). More modern plants recover a concentrated aluminum chloride solution that has found
use in municipal water treatment or industrial
oc applications. However, where demand from
such applications does not exist, disposal can
present a problem. Sometimes hazardous-waste
incineration is required.
Studies have shown ethylbenzene to be toxic
to aquatic life in relatively low concentration (10
100 mg/kg). Therefore, runoff from spills, re
control, etc. should be diked to prevent it from
entering streams or water supplies.

5. Environmental Protection

6. Quality Specications

In the United States ethylbenzene plants must


conform to the requirements of U.S. EPA National Emission Standard Organic Air Pollutants
from the Synthetic Organic Chemical Industry
(40 CFR 63, Subparts F, G, H). Producers can
comply by installing collection devices on process vents, improved seals on pumps and valves,
and oating roof tanks.
Alkylation plants that use aluminum chloride
technology produce an aqueous waste stream
from the reactor efuent wash section. In the
mid-1970s, plants produced a wet aluminum hydroxide sludge which was deposited in a landll

The product specication on ethylbenzene is set


to provide a satisfactory feedstock to the associated styrene unit. Objectionable impurities in
the ethylbenzene can be grouped into two categories: those that are detrimental to the operation
of the styrene unit and those that affect the purity of the styrene product. Impurities in product
ethylbenzene that pose an operating problem in
the conventional dehydrogenative styrene process are
1) Halides, which deactivate the dehydrogenation catalyst and contribute to downstream

10

Ethylbenzene

equipment corrosion. Usually these are chlorides originating from an AlCl3 alkylation
section or uorides from an Alkar unit.
2) Diethylbenzenes, which are dehydrogenated
to divinylbenzenes in the styrene reactor
section. The divinylbenzenes form insoluble
cross-linked polymers in the downstream process equipment. A limit of less than 10 ppm of
diethylbenzenes in the ethylbenzene product
is usually imposed.
Ethylbenzene contaminants that can affect
styrene purity are components having a boiling range between ethylbenzene and styrene.
These include xylenes, propylbenzenes, and
ethyltoluenes. The levels of cumene, n-propylbenzene, ethyltoluenes, and xylenes in the
ethylbenzene is controlled to meet the required
styrene purity specication.
A typical sales specication for a United
States manufacturer is a follows:
Purity
Benzene
Toluene
Xylenes
Nonaromatics
Propylbenzenes
Diethylbenzenes
Total chlorides as Cl
Total organic sulfur
Relative density at 15 C
APHA Color

99.5 wt % min.
0.05 0.3 wt %
0.1 0.3 wt %
0.2 wt % max.
0.05 wt % max.
0.02 wt % max.
10 mg/kg max.
1 3 mg/kg max.
4 mg/kg max.
0.869 0.872
15 max.

7. Handling, Storage, and


Transportation
Ethylbenzene is a ammable liquid. It is stored
and transported in steel containers and is subject to the control of the appropriate regulatory
agencies. The U.S. DOT identication number
is UNI 175 and the reportable quantity is 454 kg.
Details on regulations concerning the transport
of ethylbenzene can be found in the CFR or from
the DOT Material Transportation Bureau. Other
countries have regulations and safety practices
similar to those of the United States.
Foam, carbon dioxide, dry chemical, halon,
and water (fog pattern) extinguishing media are
used in ghting ethylbenzene res.
Adequate ventilation is necessary in handling
and storage areas. The use of NIOSH-approved

respirators is recommended at high concentration. Skin contact should be avoided. Chemical


gloves and safety glasses should be worn if contact is possible.
Exposure of ethylbenzene to heat, ignition
sources, and strong oxidizing agents should be
avoided.

8. Uses
Essentially all commercial ethylbenzene production is captively consumed for the manufacture of styrene monomer or in the coproduction of styrene monomer with propylene oxide. Styrene is used in the production of
polystyrene and a wide variety of other plastics
( Styrene).
Of the minor uses, the most signicant is in
the paint industry as a solvent, which accounts
for < 1 % of production capacity. Even smaller
volumes go toward the production of acetophenone, diethylbenzene, and ethylanthraquinone.

9. Economic Aspects
Ethylbenzene production is linked directly to the
styrene monomer market. A total of 99 % of
the ethylbenzene produced worldwide is used
to make styrene monomer. Through the 1960s
and into the early 1970s annual growth rates for
styrene and ethylbenzene averaged 10 %. During this period sustained growth was powered
by the expanding polystyrene market. Subsequent growth since the early 1970s has been erratic, with the price of styrene going through
four to ve year cycles, and several new ethylbenzene/styrene plants were built each time the
price started to rise. Since the mid-1980s the average growth in ethylbenzene capacity worldwide has been about 4 5 % per year, somewhat higher than the growth of the overall economy. As late as the mid 1980s most Far Eastern
ethylbenzene/styrene was produced in Japan,
and the emerging Far Eastern economies outside of Japan imported large quantities from the
West, particularly the United States and Canada.
This pattern changed in the late 1980s as capacity was added in Asia, most notably Korea,
which today holds about 10 % of the worlds

Ethylbenzene
ethylbenzene capacity. Other Far Eastern countries installing signicant new capacity have included Singapore, Thailand, and Taiwan. Over
the next ve years the expansion is likely to occur in mainland China, as multinational companies form joint ventures with local concerns to
build world-scale plants. Today each region of
the world is becoming relatively self-sufcient
in ethylbenzene/styrene.
About 15 % of ethylbenzene is consumed in
co-production of styrene and propylene oxide.
These plants offer an economic advantage because of the value of the propylene oxide coproduct, which has remained high in part because co-production technology has not been
widely available. This situation is beginning to
change, and today several producers are operating these plants in the Netherlands, Spain, Japan,
Singapore, Korea and the United States. Lower
prices for propylene oxide resulting from greater
supply could slow the use of ethylbenzene in
such plants, which have perhaps three to four
times the investment cost of an ethylbenzene dehydrogenation facility.
The 1999 worldwide capacity (in 103 t/a) by
region is as follows:
North America
South America
Western Europe
Eastern Europe
Korea/Taiwan/China
Japan
Middle East and Africa
Southeast Asia/Australia
Total

7700
330
5900
1300
3800
3500
680
500
24 700

Similar to the direct link of ethylbenzene production to that of styrene, ethylbenzene production cost is tied to feedstock cost. Modern processes all have raw material yields > 98 99 %.
Integration of the ethylbenzene and styrene processes enables efcient energy recovery of heat
from the exothermic alkylation reaction. With
90 99 % energy recovery of the heat of reaction plus the heat input to the process, production
costs are directly related to benzene and ethylene
feedstock prices. The U.S. unit sales values of
ethylbenzene in $/kg from 1960 to 1997 follow
[12]:

1960
1965
1970
1973
1974
1975
1978
1979

0.13
0.09
0.09
0.11
0.37
0.20
0.24
0.35

1980
1983
1987
1989
1990
1993
1995
1997

11

0.51
0.50
0.46
0.55
0.62
0.33
0.55
0.42

Price increases in 1974 and 1980 reect the


radical change in oil prices experienced during
these periods. Price increases in 1989/1990 and
1995 were indicative of the cyclic tightness in
supply of ethylbenzene and styrene.

10. Toxicology
The toxicology of ethylbenzene has been reviewed regularly in toxicological textbooks and
by various scientic organizations. The most recent reviews, specically relating to the exposure at the work place, are those of the German
MAK-commission [17] and of the US TLVCommittee [18] which the reader is referred to
for further details.
The acute toxicity in experimental animals
is low; oral LD50 values in rats range from 3.5
to 4.7 g/kg body weight (b.w.). A 24-h dermal
LD50 value of approximately 15 g/kg b.w. has
been reported. The acute inhalative toxicity also
is low. An 8-h exposure in experimental rats produced irritation of the respiratory tract starting at
approximately 1000 ppm, and higher concentrations induced unsteadiness, staggering gait, and
nally unconsciousness and death at 5000 ppm.
Human volunteers exposed for 7.5 h at 25 ppm
reported irritation of the mucous membranes,
which was much more pronounced at 100 ppm.
Repeated dermal applications of undiluted
ethylbenzene led to erythema, edema, and supercial necrosis. Such effects are also expected to
occur in humans due to defatting of the skin after
repeated exposures. Instillation of undiluted material into the eyes of rabbits produced a slight irritation of the conjunctival membranes but without corneal injury.
The low toxicity of ethylbenzene observed
after a single exposure was also seen in repeated dosing studies, in which signicant toxicity in experimental animals was produced only
at relatively high doses. For example, when rats

12

Ethylbenzene

and mice were exposed by inhalation over three


months no toxicity was observed in rats at 100
ppm and in mice at 500 ppm. Some increases in
lung, kidney, and liver weights were measured
starting at 250 ppm in rats and 750 ppm in mice,
but no histopathological changes were apparent
in any of the tissues. When rats and mice were
exposed to ethylbenzene by inhalation for two
years (the major part of their natural life), no effects were observed in both species at 75 ppm.
Some slight changes were found at 250 ppm, but
even at 750 ppm there were no effects on overall survival in mice and female rats. Only male
rats in the 750 ppm exposure group showed a
reduced survival as compared to untreated controls. The reduction in survival was attributed to
a rat-specic nephrotoxicity that can not be extrapolated to humans. While exposure to ethylbenzene at 750 ppm for two years had little effect
on survival, it did produce an increase in hepatic
and pulmonary tumors in mice and an increase
in the incidence of renal tumours in male rats
(see below).
Various toxicological investigations have
considered the potential reproductive toxicity
of ethylbenzene. In the three-month inhalation
study described above no effects were found on
the reproductive tissues of rats and mice at exposures up to 1000 ppm.
Other studies in which pregnant mice, rats, or
rabbits were exposed to ethylbenzene have produced divergent results. For example, increases
in fetal malformations have been reported when
pregnant mice were exposed to 113 ppm ethylbenzene. The study did not nd any other effects
indicating fetotoxicity, but insufcient information was supplied by the investigators to enable
an in-depth evaluation. Rats exposed continuously (24 h/d) up to 540 ppm showed slight maternal toxicity accompanied by fetotoxicity with
some retardations and variations of fetal development. At the same concentrations, but with
exposures lasting only 6 h/d, no adverse ndings
were noted on the dams and offspring. In another
study in which pregnant rats were exposed to
1000 ppm ethylbenzene for 6 7 h/d, maternal
toxicity was associated with slight fetotoxicity.
In an investigation with pregnant rabbits no fetal effects occurred in animals continuously exposed to ethylbenzene (24 h/d) at approximately
110 ppm. Exposure to approximately 230 ppm
was toxic to the dams, with an indication of feto-

toxicity. In contrast, no clear effects were noted


in pregnant rabbits and their offspring at 1000
ppm with a daily exposure of 7 h/d.
In vitro investigations examining the genotoxic potential of ethylbenzene have produced
conicting results. While studies in bacteria and
yeast were negative, some studies with mammalian cells in culture have produced some
weakly positive data. In contrast, in vivo investigations in drosophila and in mice [micronucleus test, test for unscheduled DNA synthesis (UDS)] were negative. In summary, while
there may be some weak potential for ethylbenzene to interact with the DNA of isolated
cells in vitro, the consistent lack of a response
in various in vivo test systems indicates that
ethylbenzene is not a mutagenic hazard. A limited in vitro mutagenicity database also is available for most of the metabolites of ethylbenzene
(see below): 1-phenylethanol, acetophenone, hydroxyacetophenone, phenyl glyoxylic acid,
mandelic acid, benzoic acid, hippuric acid, and
2-, 3-, and 4-ethylphenol. Again, although there
are some conicting results the overall weight of
evidence is that these metabolites do not exhibit
a clear genotoxic potential.
A two-year inhalation carcinogenicity study
has been carried out in rats and mice with exposures to ethylbenzene at 0, 250, and 750 ppm.
Treatment at 750 ppm produced an increased incidence of kidney tumors in male rats, an increased incidence of lung tumors in male mice,
and an increase in the incidence of liver tumors in
female mice. At 250 ppm increased tumor rates
were not observed. By comparison an oral twoyear study with rats and mice in which the animals were exposed 250 or 500 mg/kg/d1 of
mixed xylenes (containing 17% ethylbenzene)
showed no exposure-related tumor response.
As ethylbenzene does not interact with genetic material in vivo, various studies were carried out to investigate the mode of action responsible for the excess tumor formation seen in the
two-year inhalation study. In rats ethylbenzeneinduced nephrotoxicity has been shown to exacerbate chronic progressive nephropathy occurring in aging animals, which in association with
sustained cell proliferation nally results in an
increase in the background level of kidney tumors [19]. Similarly, ethylbenzene at 750 ppm
caused an increase in cell proliferation in the
lung and liver of mice, and again this prolifera-

Ethylbenzene
tive stimulus may result in excess tumor formation [20]. Further mechanistic analysis is necessary to determine whether the observed cell
replication identied in mice is likely to occur
in humans and, if so, what exposure level is relevant for establishing a health-based industrial
workplace limit. The lack of such data was the
basis for the decision of the German MAK Commission to suspend its former workplace exposure limit of 100 ppm.
Ethylbenzene is readily absorbed after inhalative, oral, and dermal exposure. Metabolism
proceeds mainly by oxidation of the side chain
and to a minor extent by aromatic-ring hydroxylation with less than 5% being excreted in urine
as 2- and 4-ethylphenol. Side-chain oxidation
mainly leads to 1-phenylethanol, one of the major urinary excretion products in rats. Further oxidative metabolism results in the production of
mandelic acid, phenyl glyoxylic acid, and benzoic acid. A second side-chain oxidative process
producing 2-phenylethanol and ultimately leading to phenyl acetic acid is a minor pathway. In
humans mandelic acid and phenyl glyoxylic acid
are the major urinary metabolites, accounting for
approximately 80% of the absorbed ethylbenzene.
Several regulatory and scientic commissions have evaluated ethylbenzene and the most
important results are summarized below.
In the EU ethylbenzene is classied and labeled with:
F and the risk phrase 11 highly
ammable
Xn and the risk phrase R 20 harmful by
inhalation
An IARC evaluation of ethylbenzene carried
out in 2000 concluded that while there was sufcient evidence for carcinogenicity in experimental animals there was inadequate evidence
for cancer in exposed humans. On this basis it
was classied as possibly carcinogenic to humans (Group 2B) [21].
The German MAK Commission formerly
had assigned a MAK value of 100 ppm. In 2001,
taking account of the carcinogenicity data, ethylbenzene was classied as a carcinogenic substance into category 3A. In the MAK process
carcinogenic chemicals are assigned to category
3A if the criteria for classication in category
4 or 5 are fullled but the database is insuf-

13

cient for the establishment of a MAK or BAT


value. For carcinogens of category 4 or 5 no
signicant contribution to human cancer risk is
to be expected . . . provided the MAK and
BAT values are observed. In other words, there
is in principle an exposure level without an increased carcinogenic risk, but the data available
currently do not allow such a specic workplace
exposure limit to be assigned. The MAK value
for ethylbenzene has therefore been suspended
while the Commission awaits the results of further mechanistic investigations to establish an
occupational exposure limit (see above) [17].
In 2002 ACGIH recommended a TLV-TWA
of 100 ppm and a TLV-STEL of 125 ppm. While
such exposure limits relate to inhalative uptake
only, the ACGIH has also dened a biological
exposure index (BEI) for mandelic acid in urine
of 1.5 g/g creatinine [22]. As mandelic acid is a
major urinary excretion product after exposure
to ethylbenzene, the advantage of a BEI is that
uptake by all exposure routes can be integrated.
With regard to the carcinogenic effect, ethylbenzene was assigned to group A3: conrmed animal carcinogen with unknown relevance to humans [18].

11. References
1. R. H. Boundy, R. F. Boyer (eds.): Styrene, Its
Polymers, Copolymers, and Derivatives,
Reinhold Publ. Co., New York 1952, p. 16.
2. American Petroleum Institute (ed.): Technical
Data Book Petroleum Rening, 12th
Revision Package to 5th ed., vol. 1, American
Petroleum Institute, Washington, D.C. 1997,
pp. 178, 179, 1112, 1113.
3. American Petroleum Institute (ed.): Technical
Data Book Petroleum Rening, metric ed.,
American Petroleum Institute, Washington,
D.C. 1981, pp. 156, 157.
4. Beilstein 5, 776 786.
5. G. A. Olah (ed.): Friedel Crafts and Related
Reactions, vol. 2, Wiley-Interscience, New
York 1964, Part 1.
6. CdF Chimie, US 4 117 023, 1978 (P. J. Gillet,
G. Henrich).
7. Monsanto, US 3 848 012, 1974 (F. Applegath,
L. E. DuPree, Jr. A. C. MacFarlane, J. D.
Robinson).
8. K. J. Fallon, H. K. H. Wang, C. R. Venkat,
UK Renery Demonstrates Ethylbenzene
Process, Oil & Gas Journal, April 17, 1995.

14

Ethylbenzene

9. Hydrocarbon Process. 76 (1997) no. 3 126.


10. Hydrocarbon Process. 78 (1999) no. 3 110.
11. CDTech, Ethylbenzene Technology, Highest
Yield Using Zeolite Based Catalyst, ABB
Lummus Global, May 1997.
12. J. Surdyk, K. L. Ring, Ethylbenzene in
Chemical Economics Handbook, SRI
International, Menlo Park, CA 1999.
13. R. J. Lewis, Sr., Saxs Dangerous Properties of
Industrial Materials, 10th ed., John Wiley &
Sons, New York, 1999.
14. ACGIH (ed.): Threshold Limit Values (TLV)
and Biological Exposure Indices, ACGIH,
Cincinnati, Ohio 1997.
15. DFG (ed.): MAK- und BAT-Werte-Liste, VCH,
Weinheim 1995.
16. F. W. Mackison, R. S. Stricoff, L. J. Partridge
(eds.): NIOSH/OSHA Occupational Health
Guidelines for Chemical Hazards, U.S. Dept.
of Health & Human Services (National
Institute for Occupational Safety & Health)

17.

18.

19.
20.

21.

22.

Publication no. 81 123, Washington, D.C.,


Jan. 1981.
H. Greim (ed.): Gesundheitsschadliche
Arbeitsstoffe;
Toxikologisch-arbeitsmedizinische
Begrundungen von MAK-Werten (Maximale
Arbeitsplatzkonzentrationen), Wiley-VCH,
Weinheim 2001.
ACGIH (ed.): Documentation of the Threshold
Limit Values for Chemical Substances, 7th ed.,
Cincinatti 2002.
G. C. Hard, Toxicol. Sci. 69 (2002) 30 41.
W. T. Stott, K. A. Johnson, R. Bahnemann, S.
J. Day, R. J. McGuirk, Toxicol. Sci. 71 (2003)
53 66.
International Agency for Research on Cancer,
IARC Monographs on the Evaluation of
Carcinogenic Risk to Humans, Vol. 77, 2000,
p. 227.
ACGIH (ed.): Documentation of the Biological
Exposure Indices, 7th ed., Cincinatti 2001.

You might also like