You are on page 1of 15

Ind. Eng. Chem. Res.

1997, 36, 1855-1869

1855

ETBE Synthesis via Reactive Distillation. 1. Steady-State


Simulation and Design Aspects
Martin G. Sneesby,*, Moses O. Tade , Ravindra Datta, and Terence N. Smith

Downloaded by UNIV NACIONAL DE COLOMBIA on August 23, 2015 | http://pubs.acs.org


Publication Date (Web): May 5, 1997 | doi: 10.1021/ie960283x

School of Chemical Engineering, Curtin University of Technology, Perth, Western Australia, and Department
of Chemical and Biochemical Engineering, University of Iowa, Iowa City, Iowa 52242

Ethyl tert-butyl ether (ETBE) is an alternative gasoline oxygenate that combines the blending
properties of methyl tert-butyl ether (MTBE) and the renewability of ethanol. Technologically,
the best means of synthesis utilizes reactive (or catalytic) distillation to maximize hydrocarbon
conversion and energy efficiency while simultaneously producing a high-purity ether product.
Mathematical models of reactive distillation are based on the conventional distillation process
with supplementary equations added to model the reactions present. Ether-alkene-alcohol
systems are highly nonideal in the liquid phase so that careful selection of physical property
routines is required to ensure satisfactory simulation results. Column simulations performed
here using both Pro/II and SpeedUp show excellent agreement with previously published
experimental data for a MTBE system and also agree well with each other for both MTBE and
ETBE systems. A homotopy analysis was performed on the tuned simulation models to
determine the effects of key design and operating variables on column performance and,
subsequently, to develop a design method for reactive distillation columns. Some unusual
behavior was identified in ETBE reactive distillation columns compared with either MTBE
columns or conventional distillation.
Introduction
Changing worldwide regulations are encouraging the
addition of oxygenates to gasoline sold in heavily
urbanized areas to reduce emissions of carbon monoxide
and unburned hydrocarbons in an attempt to combat
smog and ground level ozone. The high octane rating
of many oxygenates can also be utilized to eliminate
leaded octane enhancers, such as tetramethyllead (TML)
and tetraethyllead (TEL), from gasoline blends. To
date, methyl tert-butyl ether (MTBE) and ethanol have
been the most widely used oxygenates. MTBE appears
to offer the best combination of oxygen content, low Reid
vapor pressure (RVP), high octane, high energy content,
and low cost, but ethanol has been used in gasoline for
many years and has attracted particular interest as an
environmentally friendly alternative to fossil fuels, as
it can be produced from biomass. Many governments
also offer ethanol subsidies to offset the cost differential
with MTBE. Ethyl tert-butyl ether (ETBE) has emerged
more recently as a potential oxygenate and offers the
advantages of the blending characteristics of MTBE and
the renewability of ethanol.
Compared with MTBE, ETBE has a higher octane
rating and a lower volatility. It is also less hydrophilic
than either MTBE or ethanol and, therefore, less likely
to permeate and pollute groundwater supplies. Volatile
organic compound (VOC) emissions are also lessened
by ETBEs lower volatility compared with MTBE.
ETBE has a slightly lower oxygen content than MTBE
(and much lower than ethanol) so that larger volumes
are required, but its higher cost of production remains
its principal disadvantage when compared with either
MTBE or ethanol. However, the high cost can generally
be partially offset with renewable fuel subsidies, and
ETBE production is becoming increasingly viable in the
* Author to whom correspondence should be addressed. Telephone: 619-351-3776. Fax: 619-351-3554.
Email: sneesbym@che.curtin.edu.au.
Curtin University of Technology.
University of Iowa.
S0888-5885(96)00283-7 CCC: $14.00

current market. Table 1 (Furzer, 1994; Brockwell et al.,


1991; Lide, 1994) summarizes the key differences
between ETBE and its main alternatives. A more
complete comparison of the physical properties of ETBE
and MTBE has been published previously (Sneesby et
al., 1995).
Process schemes based on reactive distillation (also
called catalytic distillation where a catalyst is present)
are now acknowledged to offer a technological advantage
for MTBE production compared with conventional synthesis routes (Zhang et al., 1995) as yields are higher
and operating costs are lower. The high degree of
internal recycle created by the distillation operation
helps overcome the thermodynamic restriction of a
relatively low equilibrium constant and allows the
reaction to proceed further than would otherwise be
possible. The reactive distillation process is also more
energy efficient than a conventional system, as heat
generated by the exothermic reaction offsets the reboiler
heat input and contributes to product separation. High
conversion and a high ether product purity can be
obtained simultaneously in a single device. The extension of reactive distillation technology to ETBE synthesis appears to be a natural next progression.
Simulation results have been published for various
reactive distillation columns including several systems
for MTBE synthesis (Abufares and Douglas, 1995;
Sundmacher and Hoffman, 1995). However, little has
been published on ETBE synthesis. As the combination
of reaction and distillation in a single vessel can produce
interactions between various design variables which
lead to unusual responses to changes in operating
conditions, it is important to fully understand these
responses to avoid suboptimal performance and poor
designs. Furthermore, differences in phase behavior
between the MTBE and ETBE systems lead to different
sets of operating conditions and considerations, and the
simple extrapolation of concepts from MTBE synthesis
to ETBE synthesis may prove to be misleading (Sneesby
et al., 1995). Some of these issues have been addressed
1997 American Chemical Society

1856 Ind. Eng. Chem. Res., Vol. 36, No. 5, 1997

here via simulation studies using both Pro/II (Simulation Sciences, 1994) and SpeedUp (Aspen Technology,
1993).
Reaction Chemistry
ETBE is produced from the reversible reaction of
isobutylene and ethanol over an acid catalyst, such as
the acidic ion-exchange resin, Amberlyst 15:

Downloaded by UNIV NACIONAL DE COLOMBIA on August 23, 2015 | http://pubs.acs.org


Publication Date (Web): May 5, 1997 | doi: 10.1021/ie960283x

(CH3)2CdCH2 + C2H5OH S (CH3)3COC2H5(1)

(1)

The reaction is equilibrium limited in the industrially


significant range of temperatures so that the equilibrium conversion from a stoichiometric mixture of reactants at 70 C is only 84.7%. The reaction has been
studied, and detailed expressions are available for the
equilibrium constant (eq 2: Jensen and Datta, 1995)
and the reaction kinetics based on the LangmuirHinshelwood-Hougen-Watson (LHHW) model (Jensen
and Datta, 1996). Two reaction rate expressions are
givensthe generalized case (eq 3) and a simplified case
(eq 4) for conditions where the ethanol mole fraction is
greater than 4%. The proposed reaction mechanism
involves two adsorbed ethanol sites reacting with one
adsorbed isobutylene in the rate-determining step,
giving a total of three active sites. The rate equation
derived from this model produced a better fit with the
experimental data than other models.

Equilibrium constant
4060.59
T
2.89055 ln T - 0.0191544T +
5.28586 10-5T2 - 5.32977 10-8T3 (2)

Table 1. Properties of Alternative Oxygenates


property

ETBE

MTBE

ethanol

molecular weight
oxygen content (wt %)
normal boiling point (C)
blending RVP (kPa)
octane ((RON + MON)/2)
energy content (MJ/kg)
relative cost
renewable source?

102
15.7
73
27
111
36.3
moderate
partly

88
18.2
55
55
110
35.3
low
no

46
34.8
78
122
115
26.7
moderate
yes

Table 2. Key Physical Properties of the Reaction


Components
property

ethanol

isobutylene

ETBE

molecular weight
specific gravity
normal boiling point (C)
boiling point at 1000 kPa (C)
blending RVP (kPa)
specific heat (kJ/kg)
octane ((RON + MON)/2)
Energy content (MJ/kg)

46
0.795
78
155
122
2.46
115
26.7

56
0.600
-7
75
440
1.27
n/a
44.6

102
0.746
73
174
27
2.10
111
36.3

literature. An expression for the equilibrium constant


of the dimerization has previously been estimated from
the free energies of formation (eq 6: Columbo et al.,
1983).

ln K ) 95.2633 + 5819.8644/T 17.2 ln T - 0.0356T (6)


If any water is present in the reaction environment, one
further, undesirable reaction (the hydration of isobutylene to form isobutyl alcohol) is possible:

KETBE ) 10.387 +

Generalized rate equation

mcatkrateaEtOH2 aiBut rETBE )

aETBE
KETBEaEtOH

(1 + KAaEtOH)3
ln KA ) -1.0707 +

1323.1
T

krate ) 7.418 1012 exp -

(3a)

(3b)

60.4
RT

(3c)

Simplified rate equation (for EtOH > 4 mol %)

rETBE ) mcatkrate

aiBut
aETBE
2
aEtOH K
ETBEaEtOH

krate ) 1.209 1012 exp -

87.2
RT

(4a)

(4b)

The ETBE (and MTBE) reaction system also includes


an unavoidable side reactionsthe dimerization of isobutylene to produce diisobutylene (DIB):

(CH3)2CdCH2 + (CH3)2CdCH2 S [(CH3)2CdCH2]2


(5)
This reaction is also equilibrium limited, but its kinetics
and mechanism are not readily available in the open

(CH3)2CdCH2 + H2O S (CH3)3COH

(7)

It can be seen from both ETBE rate equations (eqs 3


and 4) that ethanol has a retarding effect on the
reaction. However, in practice, some ethanol excess is
required to prevent significant side reactions involving
isobutylene. The LHHW reaction model predicts a large
adsorption equilibrium constant for ethanol, which
implies that, at ethanol excesses of 4 mol % and above,
the catalyst surface is largely covered with ethanol.
Under these conditions, the dimerization and oligomerization of isobutylene are essentially eliminated (Kitchaiya and Datta, 1996).
The ETBE reaction rate also depends on the activity
of the catalyst which is susceptible to both deactivation
(slow aging) and poisoning (fast aging). Poisoning is
potentially a serious problem as water and, especially
salts, neutralize active catalyst sites. Deactivation
occurs over a much longer period and is accelerated by
thermal degradation caused by hot spots due to inadequate mixing. In situ regeneration has generally been
unsuccessful and a regular catalyst changeover is
required for most reactors (Flato and Hoffman, 1992).
The key physical properties of the reaction components differ significantly and are summarized in Table
2 (Simulation Sciences, 1994; Furzer, 1994; Brockwell
et al., 1991; CRC Handbook, 1995). In an industrial
situation, isobutylene is likely to only be available as a
component within a mixture of other, nonreactive butylenes. For most practical purposes, the physical
properties of other butylenes present can be assumed
to be similar to those of isobutylene. There is variability
in the octane data reported in the literature for each of
the various ethers, but ETBE is generally claimed to

Ind. Eng. Chem. Res., Vol. 36, No. 5, 1997 1857


Table 3. Significant Azeotropes in the ETBE Reaction
System
composition
at 0 kPa
EtOH-iBut
EtOH-nBut
EtOH-ETBE

Downloaded by UNIV NACIONAL DE COLOMBIA on August 23, 2015 | http://pubs.acs.org


Publication Date (Web): May 5, 1997 | doi: 10.1021/ie960283x

EtOH-iBut
EtOH-nBut
EtOH-ETBE

composition
at 950 kPa

composition
at 1400 kPa

Reported from Experimental Data


n/a
n/a
0.94% EtOH
n/a
n/a
n/a
37% EtOH
n/a
n/a
Predicted by UNIFAC
no azeotrope
no azeotrope
no azeotrope
no azeotrope
38% EtOH
59% EtOH

1.25% EtOH
1.45% EtOH
66% EtOH

have an average octane rating one number higher than


MTBE (Furzer, 1994; Brockwell et al., 1991; Piel and
Thomas, 1990; Unzelman, 1989). An octane rating is
not generally cited for isobutylene due to its high
volatility, but a blending value of around 100 is anticipated based on isoolefin trends.
The ETBE system is susceptible to azeotropes due to
nonidealities in the liquid phase. Azeotropes between
ethanol and isobutylene and between ethanol and ETBE
have been recorded experimentally (Gmehling et al.,
1994). The UNIFAC model predicts the presence of
these azeotropes and also suggests an azeotrope between ethanol and 1-butylene. The literature data are
incomplete, but the UNIFAC model indicates that the
azeotropes between ethanol and butylenes only exist at
high pressure. The compositions of all azeotropes vary
with pressure according to the data shown in Table 3.

ETBE Synthesis Route Using Reactive


Distillation
The conventional process for synthesis of MTBE
consists of four stages: (a) pretreatment; (b) reaction;
(c) purification; (d) recovery of unreacted reactants
(ARCO Chemical Technology, 1995). A simplified process flow diagram is shown in Figure 1. The hydrocarbon
feed should be rich in isobutylene. In practice, this
implies a refinery sourcesvarious fluidized catalytic
cracking (FCC) units (15-35% isobutylene, depending
on the type of catalyst in use), a steam cracking unit
(40-55% isobutylene), or an isobutane dehydrogenation
unit (40-55% isobutylene) (Miracca et al., 1996). The
ethanol feed should be essentially pure to minimize the
side reaction of isobutylene hydration.
The first stage of the conventional process, pretreatment, consists of a simple water wash and is necessary
to remove trace contaminants that could deactivate the
catalyst. The second stage, reaction, is conducted in two
sequential reactors to ensure that high conversions are
achieved. The first reactor normally operates isothermally and is used to perform the majority of the
reaction. A tubular reactor is normally used to allow
the substantial heat which is liberated to be removed
adequately before it affects the reaction equilibrium. An
adiabatic reactor can be used, if a slipstream of the
product is cooled and recycled to the reactor inlet to
control the temperature rise, but a lower conversion is
expected and the risk of hot spots deactivating the
catalyst is increased (Sarathy and Suffridge, 1993). The
second stage can mostly be operated adiabatically as
much less heat is liberated and a packed-bed reactor is
more economical. The packed bed allows more catalyst
to be used so that the reactor can be operated at lower
temperatures to improve the reaction equilibrium and

maximize conversion. The first reactor operates up to


90 C, while the second reactor operates at 50-60 C.
The product from the second reactor is then purified by
distillation in the third stage of the process. The
bottoms product from the ether purification column is
high-purity MTBE (>99.5%). The distillate product is
further processed to recover unreacted methanol in the
final stage. A liquid-liquid extraction column with
water as the solvent is used due to the presence of an
azeotrope between methanol and butylenes (at around
10% methanol). The raffinate product from this column
contains essentially all the nonreactive hydrocarbon
from the original hydrocarbon feed and is suitable for
alkylation as the concentration of isoolefins is very low.
One further column is required to separate water, which
is recycled to the L-L extractor, from the methanol,
which is recycled to the reactors (ARCO Chemical
Technology, 1995).
The reactive distillation process for MTBE synthesis
combines elements of the second and third stages
(reaction and purification) into one element of process
equipment, according to the simplified flow diagram
shown in Figure 2 (ARCO Chemical Technology, 1995).
The second reaction stage and peripheral equipment is
eliminated, and the overall conversion is actually increased as the effective methanol excess is much higher
within a distillation system. It is economical to maintain the first reaction stage as the amount of catalyst
that can be loaded into a reactive distillation column is
limited and the operating temperatures therein are
restricted by the intersection with separation conditions.
The recovery equipment downstream of the reactive
distillation column operates essentially unchanged, and
product purities are as high as would be achieved with
a conventional distillation operation.
Conventional ETBE synthesis follows a similar route
to MTBE synthesis and uses the same process equipment. However, reaction and phase differences create
slightly different operating conditions and change the
expected conversion and purity. The more restrictive
equilibrium lowers the conversion achieved in the
reaction stage by 1-2%. The absence of any significant
azeotrope between ethanol and butylenes results in
unreacted ethanol being recovered with the ETBE
product. The presence of an azeotrope between ethanol
and water results in some water being recycled to the
reaction stage via extract from the liquid-liquid extractor. However, there is a much lower load on the
recovery equipment, as only a small amount of ethanol
appears with the distillate product from the main
purification column. In fact, it is feasible, under some
refinery conditions, to eliminate the ethanol recovery
equipment altogether, creating the simplified flow diagram shown in Figure 3.
A reactive distillation process for ETBE synthesis
would utilize the same principles as the MTBE process
and should yield the same benefits of increased conversion, increased energy efficiency, and reduced capital
cost. The majority of the reaction (say, 80%) would be
performed in an isothermal, tubular reactor operating
at moderate conditions of temperature (around 90 C)
and pressure (1500-2000 kPa to ensure all components
remain in the liquid phase). The feed to the reactive
distillation column would, therefore, be rich in ETBE
but still contain some ethanol and isobutylene. The
products from the bottoms and overheads of the reactive
distillation column would be ETBE with some ethanol
and nonreactive hydrocarbon with a small amount of

1858 Ind. Eng. Chem. Res., Vol. 36, No. 5, 1997

Downloaded by UNIV NACIONAL DE COLOMBIA on August 23, 2015 | http://pubs.acs.org


Publication Date (Web): May 5, 1997 | doi: 10.1021/ie960283x

Figure 1. Conventional MTBE synthesis route.

Figure 2. MTBE synthesis via reactive distillation.

Figure 3. Conventional ETBE synthesis route without ethanol recycle.

isobutylene and ethanol, respectively. The distillate


product may or may not require further processing
depending on its composition and the refinery configuration. This scheme is shown in Figure 4 without
ethanol recovery equipment.
Reactive Distillation Model
Both the MESH (Material balance, vapor-liquid
Equilibria equations, mole fraction Summations, and

Heat balance) equations for systems in vapor-liquid


equilibrium and the MERQ (Material balance, Energy
balance, Rate equations for mass transfer, and phase
eQuilibrium at the vapor-liquid interface) equations for
mass transfer can be used to model distillation processes. The MERQ equations have been recommended
by some researchers (Sundmacher and Hoffman, 1995)
and are gaining in popularity due to advances in
computational power available. However, they are more
complex, require the estimation of more empirical

Ind. Eng. Chem. Res., Vol. 36, No. 5, 1997 1859

parameters, and appear to offer no improvement in


accuracy for most systems. The MESH model was used
here (see eqs 8-37) and produced satisfactory results.

Total Condenser

Separation Stage

component balance
Lin + Vin - Lout - Vout ) 0

material balance
component balance

energy balance

(8)

Linxi,in + Vinxi,in Loutxi,out - Voutxi,out ) 0 (9)

L
V
LoutHi,out
- VoutHi,out
) 0 (10)

Downloaded by UNIV NACIONAL DE COLOMBIA on August 23, 2015 | http://pubs.acs.org


Publication Date (Web): May 5, 1997 | doi: 10.1021/ie960283x

Pyi ) ixiPvap
i

(11)

TV ) TL

(12)

yi ) 1

(13)

(28)

Vyi - Lxi ) 0

(29)

VHV - LHL - Qc ) 0

(30)

equilibrium

P)

material balance

iyiPvap
i

Lin - Lout - Vt ) 0

component balance

energy balance

(31)

(32)

Linxi,in - Loutxi,out - Vyi ) 0


(33)

L
LinHin
- LoutHLout - VHV + Qr ) 0
(34)

Pyi ) ixiPvap
i

(35)

TV ) TL

(36)

(14)

yi ) 1

(37)

r1,ETBE ) -r1,EtOH ) -r1,iBut

(15)

r2,DIB ) -2r2,iBut

(16)

The basic set of MESH equations was modified for


reactive distillation by adding equations to model the
chemical reactions present. The equations shown here
are fully rigorous in that both reaction equilibrium and
reaction rate expressions are included, but chemical
equilibrium could have been assumed for this system
due to the relative speed of the reaction. This would
have been done by eliminating eqs 18, 20, and 21
(effectively assuming an infinite supply of catalyst). The
construction of the equations is such that the heat of
reaction is calculated implicitly and does not need to
be separately specified. The overhead pressure and
stage-to-stage pressure drop were fixed for these simulations but could have been modeled more rigorously
by incorporating variable relationships between holdup
and vapor-liquid flows (see part 2 of this series).
However, constant stage-to-stage pressure drop is a
common assumption and does not introduce any significant error for steady-state simulations.
The choice of physical property routines is important
due to the highly nonideal nature of the ether-alcoholalkene system. The UNIFAC model has been used
successfully to predict liquid phase activities by several
researchers including developers of reaction equilibrium
constant expressions and rate equations for the ETBE
(Jensen and Datta, 1995, 1996) and MTBE (Zhang and
Datta, 1995) reactions. It has also been found (see Table
3) to accurately represent the azeotropes compared with
experimental data (Gmehling et al., 1994). Alternatively, the UNIQUAC model (Bravo et al., 1993) or the
Wilson equation (Abufares and Douglas, 1995) can be
used to predict liquid phase activities in ether systems.
Vapor phase nonideality is less significant as etherification columns operate at only modest pressures and
any one of a range of prediction methods should be
adequate. The Soave-Redlich-Kwong (SRK) method
is widely used for the prediction of enthalpy and other
properties although alternative methods (e.g., PengRobinson) would also be acceptable for the pressure and
temperature conditions.

equilibrium

Reactive Stage
Lin + Vin - Lout - Vout +

material balance

r1,j + r2,j ) 0

r1,nBut ) r1,DIB ) r2,ETBE ) r2,EtOH ) r2,nBut ) 0

mcatkrateaEtOH2 aiBut r1,ETBE )

aETBE
KETBEaEtOH

(1 + KAaEtOH)3

(17)

(18)

4060.59
T
2.89055 ln T -0.0191544T +
5.28586 10-5T2 - 5.32977 10-8T3 (19)

ln KETBE ) 10.387 +

ln KA ) -1.0707 +

1323.1
T

krate ) 7.418 1012 exp ln KDIB ) 95.2633 +

component balance

energy balance

(20)

60.4
RT

(21)

5819.26
- 17.2 ln T - 0.0356T
T
(22)
Linxi,in + Vinxi,in - Loutxi,out Voutxi,out ) r1,j + r2,j ) 0 (23)

L
V
+ VinHi,in
LinHi,in
L
V
- VoutHi,out
) 0 (24)
LoutHi,out

equilibrium

energy balance

V-L)0

Partial Reboiler

L
V
LinHi,in
+ VinHi,in
-

equilibrium

material balance

Pyi ) ixiPvap
i

(25)

TV ) TL

(26)

yi ) 1

(27)

1860 Ind. Eng. Chem. Res., Vol. 36, No. 5, 1997

Downloaded by UNIV NACIONAL DE COLOMBIA on August 23, 2015 | http://pubs.acs.org


Publication Date (Web): May 5, 1997 | doi: 10.1021/ie960283x

Figure 4. ETBE synthesis via reactive distillation.

The results presented below for various ETBE reactive distillation columns were obtained from two different commercial simulators: Pro/II (Simulation Sciences,
1994) and SpeedUp (Aspen Technology, 1993). The Pro/
II simulation results were obtained using the built-in
reactive distillation unit operation and required only the
specification of components, reaction equilibria, feed
conditions, operating pressure, column configuration,
and two operating parameters (from reboiler and condenser duties, reflux rate/ratio, and various product
specifications) to generate simulation results. The
UNIFAC model was used for liquid phase activities. The
SRK method was used for fugacity coefficients, enthalpy, and other properties. The reaction expressions
given by eqs 2 and 6 were used to model the reactions
present.
The SpeedUp model was developed from scratch using
eqs 8-37. The UNIFAC model was again used to
predict liquid phase activities. However, an ideal vapor
phase was assumed to simplify the model. Published
Antoine coefficients (Krahenbuhl and Gmehling, 1994;
Gmehling and Onken, 1977; Dean, 1992; Reid et al.,
1987) were used to predict vapor pressures and, therefore, vapor-liquid equilibrium. Rigorous reaction kinetics were used to model the ETBE reaction (eqs 1114), while reaction equilibrium was assumed for the DIB
side reaction (eq 15). The global system of equations
for the full model contained a total of 578 variables and
504 linear and nonlinear equations. An ETBE reaction
equilibrium model, requiring six fewer equations, was
also built to test the assumption of chemical equilibrium
used for the Pro/II simulations. The isobutylene conversion was found to be only 0.2-0.3% lower for the
second case using a moderate catalyst loading. In
comparing the two process simulators, the SpeedUp
model offers a greater flexibility and the potential to
expand the model to investigate dynamic responses, but
it required significantly more time to build and develop
these models and greater prior knowledge of the system
compared with Pro/II.
Simulation Results
The laboratory column originally used to patent the
catalytic distillation process for MTBE synthesis (Smith,
1980) was used as the basis for the ETBE simulations
presented here. Smiths column contained approximately 10 ideal stages, one of which was packed with
catalyst (Abufares and Douglas, 1995). To fit with the

synthesis route shown in Figure 4, the feed is to stage


6 and consists of a 80% prereacted mixture of isobutylene and 1-butylene (40/60 split) plus ethanol at a
stoichiometric excess of 5.0%. Three reactive stages
were specified for the ETBE column so that a higher
loading of catalyst was possible to accommodate the
lower reaction rate of ETBE compared with MTBE. This
column is shown in Figure 5. The feed rate and reboiler
duty were based on a pilot scale (154 mm diameter)
ETBE reactive distillation column which has recently
been constructed at Curtin University, Western Australia. Operating conditions were adjusted slightly from
those proposed for MTBE synthesis to compensate for
differences between the two systems and improve the
overall viability of the process. A slightly higher pressure (950 kPa compared with 690 kPa in Smiths
column) was used to increase the reaction zone temperature and, thereby, improve the reaction rate and reduce
the catalyst requirement. A lower reflux ratio (5.0
compared with 10.0) was also specified to reduce the
energy requirements. The reboiler duty was optimized
with respect to isobutylene conversion. The complete
simulation input is shown in Table 4.
The key simulation results for the ETBE column, from
both Pro/II and SpeedUp, are shown in Table 5. The
estimates of conversion and purity from the two simulators are both within 1%, and the key indicator of
bottoms temperature is only 1 C divergent. The
temperature profiles and composition profiles from each
simulation are compared in Figures 6 and 7, respectively, and also match well. The most significant
discrepancies are between estimates of ethanol concentration around the middle of the stripping section and
the temperature profile near the top of the column. The
variation in predicted ethanol concentration in the
stripping section can be attributed to the assumption
of an ideal vapor phase in the SpeedUp model as this
assumption was tested using Pro/II. Vapor phase
nonideality is low where the ethanol concentration is
less than 10 mol % so that the composition and temperature profiles are generally acceptable over the rest
of the column. The differences in temperature profiles
are likely to be the result of the slightly different vapor
pressure correlations that were used (Pro/II used builtin data, while published Antoine coefficients were used
in the SpeedUp model). The lower isobutylene conversion predicted by the SpeedUp model is at least partially
attributable to the incorporation of the full reaction
kinetics in that model as a satisfactory solution to the

Downloaded by UNIV NACIONAL DE COLOMBIA on August 23, 2015 | http://pubs.acs.org


Publication Date (Web): May 5, 1997 | doi: 10.1021/ie960283x

Ind. Eng. Chem. Res., Vol. 36, No. 5, 1997 1861

Figure 5. Reactive distillation column configuration for ETBE synthesis.


Table 4. ETBE Reactive Distillation Column Simulation Input
feed conditions

column specification

temperature (C)
rate (L/min)
composition (mole basis)

30
0.76
29.1% ETBE, 9.1% ethanol, 7.3% isobutylene,
54.5% n-butylenes

composition (weight basis)

43.3% ETBE, 6.1% ethanol, 6.0% isobutylene,


44.6% n-butylenes

overall excess ethanol (mol %)

5.0

rectification stages
reaction stages
stripping stages
total stages
overhead pressure (kPa)
condenser
reflux ratio
reboiler
reboiler duty (kW)

2
3
5
10
950
total
5.0
partial
8.26

Table 5. ETBE Reactive Distillation Column Simulation Results


property

Pro/II

SpeedUp

condenser temp (C)


reaction zone temp (C)
reboiler temp (C)
isobutylene conversion (mol %)
bottoms composition (wt %)

74
75-81
159
98.3
96.1% ETBE, 2.1% ethanol, 1.7% butylenes,
0.06% DIB
97.9% n-butylenes, 0.9% isobutylene,
1.2% ethanol
6.86
8.26
0.16 (stage 3), 0.19 (stage 4), 0.15 (stage 5)

79
80-84
160
97.4
96.5% ETBE, 2.8% ethanol, 0.7% butylenes,
0.04% DIB
97.6% n-butylenes, 1.7% isobutylene,
0.7% ethanol
6.73
8.33
0.17 (stage 3), 0.18 (stage 4), 0.12 (stage 5)

distillate composition (wt %)


condenser duty (kW)
reboiler duty (kW)
reaction rates (mol/min)

full kinetic case could not be produced using Pro/II due


to difficulties in obtaining convergence. The overall
agreement between the two simulators is excellent and
is considered to be an indication of model validity.
The key blending properties of the reactive column
bottoms product can be compared with pure ETBE and
pure MTBE. These properties were estimated using the
Pro/II database and are shown in Table 6 (Simulation
Sciences, 1994). The actual values quoted for pure
ETBE and MTBE do not necessarily correspond to
experimental data but are produced on the same basis
as the column bottoms properties for comparative
purposes only. The total fuel concentration for the
ETBE column bottoms product is defined as the ethanol
concentration plus ETBE concentration. The data in

Table 6 indicates that the presence of some ethanol in


the column bottoms product has a negligible impact on
the overall properties but the small percentage of
butylenes (1.8 wt %) significantly reduces the initial
boiling point and flash point and increases the RVP of
the mixture. This is a potential disadvantage and may
necessitate modification of the column operation and/
or design to increase the ETBE product purity. Such
modifications would include additional separation stages
and may have the unfortunate side effect of reducing
the isobutylene conversion.
Model Validation and Comparison
To validate the simulation results without experimental data, Smiths MTBE column was simulated for

1862 Ind. Eng. Chem. Res., Vol. 36, No. 5, 1997


Table 6. Gasoline Blending Properties of the ETBE
Column Product

Downloaded by UNIV NACIONAL DE COLOMBIA on August 23, 2015 | http://pubs.acs.org


Publication Date (Web): May 5, 1997 | doi: 10.1021/ie960283x

Figure 6. Temperature profile in an ETBE column.

Figure 7. Composition profiles in an ETBE column.

the case described in his patent application (Smith,


1980) using both Pro/II and SpeedUp. The numbers of
rectification, reaction, and stripping stages were based
on previous estimates (Abufares and Douglas, 1995).
The equation structure and physical property routines
used were identical to those used for the ETBE column
simulations described above. Table 7 shows both sets
of simulation results for Smiths column compared with
the limited experimental data from the patent application.
The agreement between Pro/II, SpeedUp, and the
experimental results is generally excellent, especially
for the key indicators of bottoms temperature, isobutylene conversion, and MTBE purity. The most significant discrepancy between the experimental data and the
simulation models is the estimate of DIB concentration
in the bottoms product. However, both sets of simulation results presented here predict DIB concentrations
of around 1%, which is also in agreement with the
previously published simulation results, suggesting an
alternative explanation for the values reported for the
experimental system.
The acceptable accuracy of the MTBE simulations
demonstrates the adequacy of the modeling process and
validates its extension to ETBE reactive columns where
the column configuration, operating conditions, and
phase behavior are similar. Subsequently, there is a
high degree of confidence in the ETBE simulation
results presented here. However, a pilot plant has been
built, and the results from planned experiments will
permit the simulations to be checked even more rigorously.
The present study also provides a basis for comparing
the simulation packages. Pro/II is a sequential-modelbased simulation package that contains a wide range
of built-in models which allows most unit operations to
be modeled easily and accurately. The time to develop
and run simulations is generally short, and the models

property

ETBE
column
bottoms

pure
ETBE

pure
MTBE

initial boiling point (C)


final boiling point (C)
RVP (kPa)
flash point (C)
total fuel concentration (wt %)
energy content (MJ/kg)
oxygen content (wt %)
octane rating ((RON + MON)/2)
specific gravity

-6
80
41
-48
98.2
36.2
15.8
111
0.743

73
73
29
-19
100
36.3
15.7
111
0.746

55
55
55
-30
100
35.3
18.2
110
0.747

Table 7. Key Simulation Results and Experimental Data


for Smiths MTBE Column
property

Pro/II

SpeedUp

experimental

condenser temp (C)


reaction zone temp (C)
reboiler temp (C)
isobutylene conversion (mol %)
bottoms MTBE purity (mol %)
DIB bottoms concentration (mol %)
condenser duty (kW)
reboiler duty (kW)

61.0
69.0
128.0
91.6
92.2
1.2
2.14
2.32

64.6
70.1
125.7
91.6
92.6
0.9
1.99
2.16

n/a
71
127
91
91.9
6.1
n/a
n/a

are robust enough to converge when only a few initial


estimates are provided. A reactive distillation module
is present, based on a standard distillation module.
Either chemical equilibrium or kinetics can be assumed,
although the likelihood of convergence is reduced where
reaction kinetics are specified. SpeedUp is an equationbased simulator and offers the benefits of greater
flexibility, but model development is significantly longer
unless library models are used, in which case the
flexibility benefits are partially lost. Convergence criteria is much more stringent, and many more preset
variables are needed to produce a converged solution.
However, SpeedUp models can be expanded to allow the
prediction of dynamic responses and can handle reaction
kinetics with greater ease provided the equation structure is correct. For the majority of steady-state simulation objectives, Pro/II is the preferred simulator, but
SpeedUp offers potential for the study of process control
applications, startup analysis, and hysteresis phenomena.
Homotopy Analysis
Reactive distillation columns behave substantially
differently from conventional distillation columns due
to the interactions between the chemical reactions
present and vapor-liquid equilibrium. The effects of
key design and operating variables are discussed in
detail below with reference to the ETBE column studied
above. There are also substantial differences between
ETBE reactive distillation and the more common MTBE
reactive distillation, and comparisons are made where
relevant. The effects described below should be considered during design and operation of the column to
ensure optimal performance.
Effects of Feed Composition. The hydrocarbon
feed composition to an etherification unit is essentially
fixed by upstream plant operations and usually varies
between 15% and 55% reactive isobutylene depending
on the type of unit and type of catalyst in use upstream.
Standard FCC units produce C4 streams with 15-20%
isobutylene, while FCC units equipped with new generation catalysts produce C4 streams with up to 35%

Ind. Eng. Chem. Res., Vol. 36, No. 5, 1997 1863


Table 8. Effect of Hydrocarbon Feed Composition on the
Maximum Conversion and Corresponding Ether Purity
in Two ETBE Columns
10-stage column

30-stage column

isobutylene
max
ether
max
ether
concentration conversion
product
conversion
product
(mol %)
(mol %) purity (wt %) (mol %) purity (wt %)

Downloaded by UNIV NACIONAL DE COLOMBIA on August 23, 2015 | http://pubs.acs.org


Publication Date (Web): May 5, 1997 | doi: 10.1021/ie960283x

15
20
30
40
50
60

98.5
98.8
98.7
98.3
97.5
95.7

95.0
96.2
96.7
96.1
95.4
94.1

98.7
98.6
94.5
91.7
88.5
86.6

97.2
96.9
94.8
93.6
91.6
91.1

isobutylene. The C4 streams from steam cracking and


isobutane dehydrogenation units are even richer in
isobutylene (up to 55%) (Miracca et al., 1996). In each
case, the other components are predominantly n-butylenes although other C4s (mainly isobutane and nbutane) are also present in some cases. Although the
operation of the various catalytic and steam cracking
units can be varied to increase or decrease isobutylene
production, other factors (e.g., maximizing gasoline
production) usually have a more significant economic
impact and govern the operation of the unit.
Increasing the concentration of reactive isobutylene
in the hydrocarbon feed has four main effects on reactive
distillation column operation: (a) the energy cost, per
kg of ETBE produced, decreases as less energy is
wasted heating and cooling inert components; (b) the
reactant concentrations in the reaction zone increase,
with a favorable effect on the reaction equilibrium; (c)
the reaction zone temperatures and the reaction zone
temperature gradient increase as the stabilizing effect
of inert components is lessened, with a detrimental
effect on the reaction equilibrium; (d) the specific
reboiler duty must be decreased to maintain optimal
reaction conditions, with a detrimental effect on product
purity. The overall effect of increasing isobutylene in
the feed is usually to decrease the maximum isobutylene
conversion and corresponding product purity.
The effect of isobutylene concentration on reactive
distillation column operation contrasts with MTBE
synthesis where an isobutylene feed concentration of
around 60% is optimal for conversion and energy
efficiency. A consequence of this result is that, when
the isobutylene concentration in the hydrocarbon feed
is low, ETBE synthesis may be more favorable than
MTBE synthesis for some column configurations. However, the presence of a significant azeotrope in the
MTBE system, between methanol and butylenes, means
that isobutylene conversion and MTBE product purity
can essentially be increased as far as is desired by
adding more stages and/or increasing reflux. This does
not necessarily apply to ETBE systems (see discussion
in the Effects of Separation Stages section).
The maximum conversion in a 10-stage ETBE reactive distillation column (Figure 5) and a 30-stage ETBE
reactive distillation column based on a commercial
MTBE column (Simulation Sciences, 1995) (where the
co-objective is to essentially eliminate butylenes from
the ether product) was determined for varying isobutylene concentrations in the hydrocarbon feed to the
primary reactor, by simulations using Pro/II. For all
cases, 80% conversion in the reactor was assumed and
the reactive distillation reboiler duty was optimized with
respect to the final, overall conversion. Table 8 shows
the maximum final conversions and the corresponding
ETBE product purities.

ETBE synthesis from pure isobutylene and pure


ethanol (three-component system only) is not feasible,
as the intersection of phase and chemical equilibrium
is not favorable. Too little isobutylene is available in
the liquid phase for reaction, and, at pressures that
result in favorable reaction equilibrium, the separation
of ethanol and ETBE is difficult. The maximum isobutylene conversion and ether purity attainable in a
ternary system from the column configuration shown
in Figure 5 are around 83 mol % and 91 wt %,
respectively.
Effects of Stoichiometric Excess of Ethanol. The
stoichiometric ratio of reactants significantly affects the
reaction conversion and the loads on the downstream
product purification and reactant recovery equipment.
If the reactant excess is too low, product conversion is
adversely limited, while if it is too high, purification
costs are increased and/or product purity is decreased.
The choice of percent excess essentially becomes a
compromise between operating costs and the value
added by the process (a function of market conditions).
The control of the stoichiometric ratio is often complicated by the inability to accurately measure feed
composition and the need to ensure that certain constraints (resulting from reaction kinetics or other sources)
are never violated.
In an MTBE column, the majority of unreacted
methanol is recovered in the distillate product via
azeotropes that occur between methanol and butylenes.
This places an upper limit on the methanol excess as
sufficient butylenes must always be present in the
distillate product to maintain the unreacted methanol
below the azeotropic composition (around 10 mol %
methanol). For example, if the hydrocarbon feed contained 60% isobutylene and 40% n-butylenes and the
azeotropic composition at the operating pressure was
10 mol % methanol, then the maximum methanol excess
would be approximately given by:

unreacted MeOH e 10% (nBut + unreacted


iBut)
e 10% (41% total
hydrocarbon feed)
e 4.1% (total hydrocarbon feed)
reacted MeOH ) reacted iBut 59% (total
hydrocarbon feed)
maximum MeOH excess (4.1%/59%) ) 6.9%
Another consequence of recovering methanol overhead
via azeotropes is that, below the maximum methanol
excess as determined by the feed and azeotropic compositions, increasing the methanol excess has only a
limited effect on the bottoms product purity.
However, in an ETBE column, unreacted ethanol is
recovered directly with the ether product. This removes
the restriction on ethanol excesses that can be used but
increases the effect that the ethanol excess has on the
ether purity. Therefore, a compromise must be determined between isobutylene conversion (which rises as
the ethanol excess increases) and ether purity (which
falls as the ethanol excess increases). The column
configuration influences this decision by changing the
relative magnitudes of the two effects, but an ethanol
excess between 4.0 and 7.0 mol % is considered sufficient to produce high isobutylene conversions without

1864 Ind. Eng. Chem. Res., Vol. 36, No. 5, 1997


Table 9. Effect of Column Overhead Pressure on the Maximum Isobutylene Conversion and Corresponding Ether
Purity in Two ETBE Columns

Downloaded by UNIV NACIONAL DE COLOMBIA on August 23, 2015 | http://pubs.acs.org


Publication Date (Web): May 5, 1997 | doi: 10.1021/ie960283x

10 stage-column

30 stage-column

overhead
pressure (kPa)

reaction zone
temp (C)

maximum
conversion (mol %)

ether purity
(wt %)

reaction zone
temp (C)

maximum
conversion (mol %)

ether purity
(wt %)

400
500
600
700
800
900
1000
1100

44-53
51-59
57-65
63-70
68-74
72-79
77-82
81-86

99.0
99.2
99.1
99.0
98.8
98.5
98.2
97.8

97.0
97.1
97.1
97.0
96.6
96.4
95.9
95.3

44-50
51-56
57-62
62-67
67-72
72-77
77-81
81-85

99.1
98.9
98.7
98.6
98.6
98.2
98.0
97.7

97.1
97.1
97.3
97.6
98.1
98.3
99.1
99.9

Figure 8. Effects of the stoichiometric excess of ethanol on


isobutylene conversion. Ether purity and distillate composition
from an ETBE column.

Figure 9. Effects of reactive stages on isobutylene conversion


from an ETBE column.

overly diluting the ether product with ethanol (thereby


adding to downstream recovery costs) while providing
a satisfactory driving force for the reaction. A very highpurity ether product can be produced with a lower
ethanol excess, but some excess of ethanol should
always be used to suppress side reactions involving
isobutylene. Reaction zone conditions and distillate
composition are essentially independent of the percent
excess of ethanol in an ETBE column. Using the column
configuration shown in Figure 5 as the basis for simulations, Figure 8 shows the effect of increasing the ethanol
excess on isobutylene conversion and ETBE purity.
Effects of Column Pressure. In conventional
distillation, the operating pressure of a column is
normally set through an economic rationalization of
heat-transfer costs and the value of improved separation
(via increasing relative volatility with reducing pressure) (Kister, 1992). However, in reactive distillation,
the choice of operating pressure is complicated by the
indirect effect of pressure on the reaction equilibrium
via changing phase equilibrium temperaturessincreasing
the pressure raises the reaction zone temperature and

Figure 10. Effects of rectification separation on isobutylene


conversion and distillate composition from an ETBE column.

Figure 11. Effects of stripping separation on isobutylene conversion and ether purity from an ETBE column.

decreases the reaction equilibrium constant of exothermic reactions such as ETBE synthesis, thereby lowering
conversion. The effect of pressure on the rate constant,
via VLE temperature changes, must also be considered,
if the reaction is kinetically controlled.
For both ETBE and MTBE systems, the nonideality
of the liquid phase adds two further restrictions to
column operations. First, below a certain pressure
(about 130 kPa for ETBE and 300 kPa for MTBE
columns) the highest boiling component is the alcohol
rather than the ether. Operating below this pressure
will drive the ether product back to the reaction zone
and will result in low conversion and low ether purity.
Second, the overhead pressure affects the composition
(and presence) of the alcohol-butylene azeotropes. This
is of greater significance for MTBE columns as the
azeotrope is more common and contains a larger fraction
of unreacted alcohol but should not be ignored for ETBE
columns.
The range of effects that need to be considered in
selecting the column operating pressure suggests that
accurate simulation results are almost essential for
reactive distillation column design. Table 9 shows the

Ind. Eng. Chem. Res., Vol. 36, No. 5, 1997 1865


Table 10. Effect of Separation Stages on the Maximum Conversion and Corresponding Ether Purity in ETBE Columns
with Lean Isobutylene Feeds
rectification
stages
2
4
8

stripping stages
8

16

98.6 mol % conversion;


90.3 wt % purity
(8.8 wt % butylenes)
98.8 mol % conversion;
92.2 wt % purity
(6.8 wt % butylenes)
99.0 mol % conversion;
98.6 wt % purity
(0.16 wt % butylenes)

99.4 mol % conversion;


99.7 wt % purity
(0.21 wt % butylenes)
99.5 mol % conversion;
99.2 wt % purity
(0.17 wt % butylenes)
99.5 mol % conversion;
98.6 wt % purity
(0.16 wt % butylenes)

99.4 mol % conversion;


99.95 wt % purity
(0 wt % butylenes)
99.5 mol % conversion;
99.7 wt % purity
(0 wt % butylenes)
99.5 mol % conversion;
98.9 wt % purity
(0 wt % butylenes)

Table 11. Effect of Separation Stages on the Maximum Conversion and Corresponding Ether Purity in ETBE Columns
with Rich Isobutylene Feeds
rectification
stages

Downloaded by UNIV NACIONAL DE COLOMBIA on August 23, 2015 | http://pubs.acs.org


Publication Date (Web): May 5, 1997 | doi: 10.1021/ie960283x

2
4
8

stripping stages
8

16

97.3 mol % conversion;


93.1 wt % purity
(6.6 wt % butylenes)
97.2 mol % conversion;
93.2 wt % purity
(3.6 wt % butylenes)
97.1 mol % conversion;
93.2 wt % purity
(3.5 wt % butylenes)

97.8 mol % conversion;


97.0 wt % purity
(0.05 wt % butylenes)
98.2 mol % conversion;
97.0 wt % purity
(0.08 wt % butylenes)
98.0 mol % conversion;
96.7 wt % purity
(0.14 wt % butylenes)

98.2 mol % conversion;


97.2 wt % purity
(0 wt % butylenes)
98.2 mol % conversion;
97.1 wt % purity
(0 wt % butylenes)
98.2 mol % conversion;
97.1 wt % purity
(0 wt % butylenes)

maximum conversion attainable in two ETBE columns


for varying overhead pressures, assuming chemical
equilibrium is attained. The first column is the base
case shown in Figure 5, while the second column
corresponds to the first row of Table 8 (30 stages and
15% isobutylene in the feed). The optimum with respect
to conversion is 400-500 kPa, although higher pressures are preferred for kinetically controlled reactions
as the reaction rate at temperatures corresponding to
this pressure are very low. The optimum with respect
to ETBE purity occurs at a significantly higher pressure
in the 30-stage column, and very high purities (greater
than 99.9 wt %) are possible at high pressures (greater
than 1100 kPa). This suggests that an intermediate
pressure (e.g., 700-900 kPa) is favored overall. The
range of effects present prevents the recommendation
of a standard operating pressure for ETBE columns,
although an optimum will usually exist. Site factors
and the expected variation in feed composition are likely
to be decisive in optimizing the operating pressure.
However, as a rule-of-thumb, it is considered that
operating at a column pressure such that the reaction
zone temperatures are 5-15 C lower than those used
in the prereactor should guarantee high conversion,
high ether product purity, and manageable reaction
rates.
Effects of Reactive Stages. The function of the
reactive section of the column is simply to provide a site
for the main reaction to proceed, and, as such, there is
no particular requirement for separation. This suggests
that only one equilibrium stage of a column needs
catalyst present, although the actual physical size of this
stage could be quite large to meet catalyst requirements.
However, simulations show that higher conversions are
possible where more than one equilibrium stage is
reactive. Figure 9 shows simulation results for an
ETBE column where the number of reactive stages was
varied. All other variables, including the number of
separation stages, reflux ratio, reboiler duty, and feed
conditions, were unchanged.
The improved conversion with an increased number
of reactive stages results from the benefits of the

Figure 12. Effects of reflux ratio on isobutylene conversion. Ether


purity and energy requirement of an ETBE column.

Figure 13. Effects of reboiler duty on isobutylene conversion and


ether purity from an ETBE column.

additional separation which is gained. Under most


conditions, transferring all the catalyst in a column to
a single stage would have a negligible effect on the
overall conversion achieved. Note that this is different
from the data presented in Figure 9 as it implies an
increase in the number of separation stages at the
expense of reactive stages.
Increasing the number of reactive stages above the
optimum (four for this column) produced a detrimental
interaction between the phase and chemical equilibrium

1866 Ind. Eng. Chem. Res., Vol. 36, No. 5, 1997

Downloaded by UNIV NACIONAL DE COLOMBIA on August 23, 2015 | http://pubs.acs.org


Publication Date (Web): May 5, 1997 | doi: 10.1021/ie960283x

Table 12. Suggested Reactive Distillation Design Strategy for Ether Oxygenates
1. Define values for the key process objectives of hydrocarbon conversion and ether product purity and determine
the composition of the hydrocarbon feed stream.
2. Identify the key components in the rectification and stripping sections and the effects on the composition profiles of
increasing rectification and stripping separation for the specific feed composition to be used.
3. Select a value for the stoichiometric excess of alcohol reactant considering the conversion and purity requirements identified above
(high for maximum conversion; low for maximum purity).
4. Estimate operating pressure based on reaction equilibrium constant at the boiling point of a mixture that is predominantly
hydrocarbon with some ether and alcohol.
5. Perform rigorous simulations with a range of numbers of rectification and stripping stages and a range of reflux ratios.
6. The minimum reflux ratio cannot easily be determined for reactive distillation because of interactions with chemical equilibrium.
Select a combination of rectification stages, stripping stages, and reflux ratio based on process objectives, local utility costs,
and product values. Note that, for some feed compositions and stoichiometric excesses, the rectification and stripping separation
must be optimized to prevent excessive loss of unreacted reactants in both products and that for these cases there
exists a limit to the conversion and ether purity achievable.
7. Where feed rates are relatively low, ensure the reflux ratio selected is sufficient to adequately load a column of 1.2 m
or greater diameter.
8. Determine the number of reactive stages required using simulations with the selected reflux ratio and number of stages.
Add reactive stages until decomposition occurs on the lowest reactive stages and/or there is no incremental benefit
to isobutylene conversion.
9. Select a value of stage efficiency based on conventional distillation and determine the number of actual stages required
for the rectification and stripping sections.
10. Estimate reboiler and condenser duties from simulations based on the above assumptions.
11. Determine column diameter from simulation data for vapor and liquid loadings and column height from stage efficiency
estimates, including appropriate allowances for uncertainties in flooding factor and stage efficiency.
12. If required, adjust the reflux ratio to vary column width and height to produce an acceptable design. Note the impact of
reducing reflux ratio on conversion over and above its effect on separation and the need to optimize separation for some cases.
13. Based on the column configuration defined above, optimize stoichiometric excess, pressure,
and reboiler duty to maximize hydrocarbon conversion and ether product purity using further simulations.

which led to the decomposition of product on the lower


reactive stages. This is to be expected as the extra
separation occurring with more stages concentrates the
ether in the lower reactive stages and shifts the chemical equilibrium back to the reactants. An excessive
number of reactive stages can also encourage unwanted
side reactions and increase the concentration of impurities in the ether product. By comparison, the MTBE
case showed no optimum although the benefits of adding
a reactive stage became progressively smaller when
high numbers of reactive stages were already present,
due to the increased likelihood of decomposition on the
lower reactive stages.
Supplying catalyst on several reactive stages allows
the total catalyst loading in the column to be increased
and may, therefore, extend the time between catalyst
changeovers/regenerations. However, during the life of
the catalyst the main reaction site may shift within the
column and change the effective number of rectification
and stripping stages and, subsequently, change the
conversion and purity attained.
Effects of Separation Stages. Ideally, the rectification section of a reactive distillation column for ether
synthesis should (a) remove light inerts from the reaction zone, (b) prevent loss of ether or alcohol in the
distillate, and (c) recycle unreacted reactants (olefin and
alcohol) to the reaction zone. For an ETBE column, this
would ideally require a separation between isobutylene
and the heaviest nonreactive hydrocarbon lighter than
isobutylene. In practice, this is almost impossible to
achieve while maintaining acceptable reaction zone
conditions. However, the loss of ether in the distillate
can be minimized without rejecting isobutylene from the
column. More rectification stages are required to also
prevent loss of ethanol with the distillate.
Ideally, the stripping section should (a) remove ether
from the reaction zone to maintain favorable reaction
conditions, (b) purify the ether product, (c) prevent loss
of reactive olefin with the ether product, and (d)
minimize ethanol loss with the ether product. In an
ETBE column, this implies a separation between ethanol and ETBE, which is largely achievable at the

conditions of temperature and pressure required for


adequate reaction. Manipulating the number of stripping stages provides a mechanism for controlling the
volatility, flash point, and composition of the ether
product.
Although the separation objectives are clear, in a
multicomponent system changes in the separation efficiency tend to adjust composition profiles rather than
produce more clearly defined product splits when intermediate boiling components (e.g., ethanol) are involved. Consequently, an increase in rectification or
stripping separation is not necessarily beneficial for
reaction zone conditions. Too much rectification separation can result in isobutylene loss via the distillate.
Too much stripping separation can result in ethanol
being drawn away from the reaction zone and concentrated just above the reboiler. For some column configurations, both rectification and stripping separation
must be optimized. Figures 10 and 11 show examples
of this phenomenon for a column based on the configuration given in Figure 5.
The ratio between the number of rectification and
stripping stages and the feed composition are also
important factors that should be considered. If the
rectification separation is too great without light inerts
present to stabilize the reaction conditions, then reboiler
operation must be adjusted to compensate the reaction
conditions and, subsequently, the purity of the ether
product is sacrificed. Tables 10 and 11 show this effect
for two different feed compositions: a low isobutylene
feed with 15% isobutylene in the hydrocarbon to the
primary reactor and a high isobutylene feed derived
from a 50/50 mixture of isobutylene and n-butylenes.
The number of reactive stages and the reflux ratio were
kept constant for all simulations, and the reboiler duty
was optimized with respect to conversion in each case.
The best combination of rectification and stripping
stages is 2 and 16, respectively. Adding rectification
stages is beneficial with few stripping stages present
but detrimental with many. With a high concentration
of isobutylene in the hydrocarbon feed, the magnitudes

Downloaded by UNIV NACIONAL DE COLOMBIA on August 23, 2015 | http://pubs.acs.org


Publication Date (Web): May 5, 1997 | doi: 10.1021/ie960283x

Ind. Eng. Chem. Res., Vol. 36, No. 5, 1997 1867

of all effects are diminished and the need to optimize


(rather than maximize) separation becomes more likely.
Effects of Reflux Ratio. In a reactive distillation
column, reflux not only enhances separation but recycles
unreacted reactants to the reaction zone and increases
conversion. In an ETBE column, even where the
concentration of isobutylene is high and rectification
separation must be optimized to maximize conversion,
increasing reflux still increases conversion. Therefore,
separation stages and reflux cannot necessarily be used
interchangeably, and high reflux ratios may sometimes
be preferred. Several effects occur as a result of
increasing the reflux ratio: (a) the concentration of
reactants in the distillate is reduced; (b) the reaction
zone temperatures are reduced; (c) the concentration of
ether in the reaction zone is reduced. Each of these
effects contributes to increased conversion. Figure 12
indicates the relationship between reflux ratio and
conversion for the reactive distillation column shown
in Figure 5. Although conversion increases monotonically with reflux ratio, the ether product purity remains
approximately constant over a wide range of reflux
ratios due to the increasing load on the stripping section
as the amount of ETBE in the column increases.
In an industrial environment, a high reflux ratio is
economically unattractive as it adds to the equipment
size and energy requirements. In conventional distillation, reflux and stages can be traded off against each
other. In reactive distillation, this can also be done but
may incur a conversion penalty. ETBE columns are
susceptible to this penalty, and, depending on the feed
composition, the maximum conversion could be achieved
in a shorter column with high reflux. No significant
conversion penalty exists for MTBE columns, and the
most viable design is a tall column with a low reflux
ratio.
Effects of Reboiler Duty. The reboiler duty is one
of the principal points of control in a distillation column.
In normal distillation, there is a monotonic, albeit
sometimes highly nonlinear, relationship between the
reboiler duty and the principal operating objective
(Kister, 1992). In reactive distillation, the reboiler duty
must be set to ensure sufficient recycle of unreacted,
heavy reactant to the reaction zone without excluding
the light reactant from the reaction zone. If the reboiler
duty is too high or too low, conversion, and subsequently
product purity, is reduced. Figure 13 indicates the effect
of reboiler duty on conversion and ether purity and
clearly shows the presence of an optimum duty.
Effects of Other Operating and Design Variables. The extent of reaction performed prior to the
reactive distillation column should simply be an economic optimization between the relative costs of the two
operations. Initially, it is easy to produce ether in a
simple tubular reactor with some form of temperature
control to prevent the exothermicity of the reaction from
heating the reactor contents and, eventually, stopping
the reaction due to equilibrium considerations. However, at higher conversions, very low reaction temperatures are required with a subsequent high demand for
catalyst due to the low reaction rates. Under these
conditions, reactive distillation becomes more economical.
The optimum feed point to the reactive distillation
column is just below the reactive section to avoid the
possibility of ETBE decomposition with the relatively
ETBE-rich product from the reactor. Feeding too far
below the reactive section reduces the stripping poten-

Figure 14. Flowchart for reactive distillation design.

tial of the column and increases the energy required for


separation. Split feeding to each of the reactive stages
is possible but creates a high concentration of ETBE at
the top of the reactive section (leading to decomposition
toward the bottoms of the reactive section) and, again,
increases the energy required for separation.
The feed temperature has only a very slight effect on
the operation of either an ETBE or MTBE reactive
distillation column. A cooler feed has a mildly beneficial
effect on the reaction zone temperature, but this can
be offset by a shift in phase equilibrium. To minimize
equipment requirements, the feed should be supplied
at its process temperature, which is likely to be close to
the reactor temperature (80-90 C). If intermediate
storage is required for any reason, an ambient feed is
also acceptable. Neither feed heaters or feed coolers are
required for a satisfactory process design.
Process Design Method
Simulations should form a key part of reactive distillation process design, more so than in conventional

1868 Ind. Eng. Chem. Res., Vol. 36, No. 5, 1997

Downloaded by UNIV NACIONAL DE COLOMBIA on August 23, 2015 | http://pubs.acs.org


Publication Date (Web): May 5, 1997 | doi: 10.1021/ie960283x

distillation design, due to the increased complexity of


operation and the larger number of design and operating
variables. The importance of simulations is further
enhanced by the absence of satisfactory shortcut or
empirical methods that define the significant effects of
key variables. Residue curves and tie-lines (Ung and
Doherty, 1995), which parallel conventional distillation
design methods, are accepted design tools for the
reactive distillation of ternary systems but are only
applicable to the reactive section of the column. Little
has been published on integrated design methods for
reactive distillation that recognize the interactions
between all three sections (rectification, reaction, and
stripping) of a reactive distillation column. A possible
design strategy is proposed in Table 12. It is important
to note that the design process is iterative, and a
successful design may require several iterations. Figure
14 shows the same design process diagrammatically and
indicates where the process may become iterative by
showing recycle steps in dashed lines.
Conclusions
Process simulation of ETBE reactive distillation
columns can be performed using either the MESH or
MERQ distillation equations, with appropriate additional equations to model the chemical reaction(s). The
MESH method was shown to be accurate for a wellknown MTBE column and was extended to ETBE
columns using both Pro/II and SpeedUp. The results
from the two simulators were in good agreement and
were considered to be a satisfactory basis for a homotopic study of the effects of key operating variables in
ETBE reactive distillation columns.
Several useful results for ETBE synthesis via reactive
distillation were obtained from the homotopic study of
simulation columns:
(1) Ethanol is predominantly recovered with the
bottoms product in an ETBE reactive distillation column, whereas methanol is recovered with the distillate
product from a MTBE reactive distillation column.
(2) ETBE production is more attractive with low
isobutylene feed concentrations, whereas MTBE production is best with around 60% isobutylene in the feed.
(3) The numbers of rectification, reaction, and stripping section stages generally need to be optimized for
maximum isobutylene conversion.
(4) A high reflux rate is necessary for maximum
isobutylene conversion, although satisfactory operation
can be achieved with a low reflux rate and more stages.
(5) The reboiler duty needs to be tightly controlled
around a narrow optimum for satisfactory column
operation.
The results of the study were used to develop an
integrated design method for reactive distillation columns for ether synthesis. The design method relies on
accurate simulation results and an iterative process to
locate an acceptable design meeting all the process
requirements.
Nomenclature
ai ) activity of component i
HL ) molar liquid enthalpy
HV ) molar vapor enthalpy
krate ) reaction rate constant
KA ) ethanol adsorption equilibrium constant
KDIB ) DIB reaction equilibrium constant
KETBE ) ETBE reaction equilibrium constant

L ) molar liquid flow


mcat ) mass of catalyst
P ) pressure
Pvap ) vapor pressure
Qc ) condenser duty
Qr ) reboiler duty
r1,i ) reaction rate of component i in the first reaction
(ETBE)
r2,i ) reaction rate of component i in the second reaction
(DIB)
TL ) liquid temperature
TV ) vapor temperature
V ) molar vapor flow
xi ) molar liquid concentration of component i
yi ) molar vapour concentration of component i
i ) activity coefficient of component i
Abbreviations
DIB ) diisobutylene
ETBE ) ethyl tert-butyl ether
EtOH ) ethanol
iBut ) isobutylene
MON ) motor octane number
MTBE ) methyl tert-butyl ether
nBut ) normal butylenes (1-butylene and 2-butylene)
RON ) research octane number
RVP ) Reid vapor pressure

Literature Cited
Abufares, A. A.; Douglas, P. L. Mathematical Modeling and
Simulation of an MTBE Catalytic Distillation Process Using
SpeedUp and Aspen Plus. Chem. Eng. Res. Des. 1995, 73, 3.
ARCO Chemical Technology. Ethers, Petrochemical Processes 95.
Hydrocarbon Process. 1995, 74 (3), 110.
Aspen Technology Inc. The SpeedUp Users Manual; Aspen
Technology Inc.: Cambridge, MA, 1993.
Bravo, J. L.; Pyhalahti, A.; Jarvelin, H. Investigations in a
Catalytic Distillation Pilot Plant: Vapor/Liquid Equilibrium,
Kinetics, and Mass-Transfer Issues. Ind. Eng. Chem. Res. 1993,
32, 2220.
Brockwell, H. L.; Sarathy, P. R.; Trotta, R. Synthesize Ethers.
Hydrocarbon Process. 1991, 70 (9), 133.
Columbo, F.; Corl, L.; Dalloro, L.; Delogu, P. Equilibrium Constant
for the Methyl tert-Butyl Ether Liquid Phase Synthesis by Use
of UNIFAC. Ind. Eng. Chem. Fundam. 1983, 22, 219.
Dean, J. A. Langes Handbook of Chemistry, 14th ed.; McGrawHill: New York, 1992.
Flato, J.; Hoffman, U. Development and Start-up of a Fixed Bed
reaction Column for Manufacturing Antiknock Enhancer MTBE.
Chem. Eng. Technol. 1992, 15, 193.
Furzer, I. Australian Gasoline and MTBE. Chem. Eng. Aust. 1994,
19 (4), 9.
Gmehling, J.; Onken, V. Vapor-liquid Equilibrium Data Collection; DECHEMA: Frankfurt, Germany, 1977.
Gmehling, J.; Menke, J.; Krafczyk, J.; Fischer, K. Azeotropic Data,
Part 1; VCH: Weinheim, Germany, 1994.
Jensen, K. L.; Datta, R. Ethers from Ethanol. 1. Equilibrium
Thermodynamic Analysis of the Liquid Phase Ethyl tert-Butyl
Ether Reaction. Ind. Eng. Chem. Res. 1995, 34, 392.
Jensen, K. L.; Datta, R. Ethers from Ethanol. 7. Transition State
Theory Analysis of the Kinetics of Liquid Phase Ethyl tert-Butyl
Ether Synthesis Reaction. Submitted to Ind. Eng. Chem. Res.
1996.
Kister, H. Z. Distillation Design; McGraw-Hill: New York, 1992.
Kitchaiya, P.; Datta, R. Ethers from Ethanol. 6. Kinetics of
Simultaneous tert-Amyl Ethyl Ether Synthesis and Isoamylene
Isomerisation. Submitted to Ind. Eng. Chem. Res. 1996.
Krahenbuhl, M. A.; Gmehling, J. Vapor Pressures of Methyl tertButyl Ether, Ethyl tert-Butyl Ether, Isopropyl tert-Butyl Ether,
tert-Amyl Methyl Ether and tert-Amyl Ethyl Ether. J. Chem.
Eng. Data 1994, 39, 759.
Lide, D. R., Ed. CRC Handbook of Chemistry and Physics, 75th
ed.; CRC: Boca Raton, FL, 1994.
Miracca, I.; Tagliabue, L.; Trotta, R. Multitubular Reactors for
Etherifications. Chem. Eng. Sci. 1996, 51 (10), 2349.

Ind. Eng. Chem. Res., Vol. 36, No. 5, 1997 1869

Downloaded by UNIV NACIONAL DE COLOMBIA on August 23, 2015 | http://pubs.acs.org


Publication Date (Web): May 5, 1997 | doi: 10.1021/ie960283x

Piel, W. J.; Thomas, R. X. Oxygenates for Reformulated Gasoline.


Hydrocarbon Process. 1990, 69 (7), 68.
Reid, R. C.; Prausnitz, J. M.; Poling, B. E. The Properties of Gases
and Liquids, 4th ed.; McGraw-Hill: New York, 1987.
Sarathy, P. R.; Suffridge, G. S. Etherify Field Butanes (Part 2).
Hydrocarbon Process. 1993, 72 (2), 43.
Simulation Sciences Inc. Pro/II Keyword Input Manual; Simulation Sciences Inc.: Brea, CA, 1994.
Simulation Sciences Inc. Pro/II Casebook. Methyl Tertiary Butyl
Ether (MTBE) Plant; Simulation Sciences Inc.: Brea, CA, 1995.
Smith, L. A. Catalytic Distillation Process and Catalyst. Eur. Pat.
Appl. EP8860, 1980.
Sneesby, M. G.; Tade, M. O.; Datta, R. tert-Butyl EtherssA
Comparison of Properties, Synthesis Techniques and Operating
Conditions for High Conversions. Dev. Chem. Eng. Miner.
Process. 1995, 3, 89.
Sundmacher, K.; Hoffmann, U. Oscillatory Vapor-Liquid Transport Phenomena in a Packed Bed Reactive Distillation Column
for Fuel Ether Production. Chem. Eng. J. 1995, 57, 219.
Ung, S.; Doherty, M. F. Synthesis of Reactive Distillation Systems
with Multiple Equilibrium Chemical Reactions. Ind. Eng. Chem.
Res. 1995, 34, 2555.

Unzelman, G. H. Ethers in Gasolines1. Ethers Have Good


Gasoline Blending Attributes. Oil Gas J. 1989, Apr 10, 33.
Zhang, T.; Datta, R. Integral Analysis of Methyl tert-Butyl Ether
Synthesis Kinetics. Ind. Eng. Chem. Res. 1995, 34, 730.
Zhang, T.; Kitchaiya, P.; Jensen, K. L.; Phillips, C.; Datta, R.
Ethers from Renewable/Recoverable Resources. Final Technical
Report NREL/#XAT-3-13413-01; University of Iowa: Iowa City,
IA, 1995.

Received for review May 20, 1996


Revised manuscript received January 15, 1997
Accepted January 16, 1997X
IE960283X

X
Abstract published in Advance ACS Abstracts, February
15, 1997.

You might also like