You are on page 1of 14

Article No : c22_c01

Reactive Distillation
MICHAEL SAKUTH, Sasol Solvents Germany GmbH, Moers, Germany
DIETER REUSCH, Degussa AG, Marl, Germany
RALF JANOWSKY, Degussa AG, Mobile, Alabama, United States

1.
2.
2.1.
2.2.
3.
3.1.
3.2.

Introduction. . . . . . . . . . . . . . . . . . . . . . . .
Mathematical Modeling of Reactive
Distillation Processes . . . . . . . . . . . . . . . . .
Equilibrium-Based Models . . . . . . . . . . . .
Rate-Based Models . . . . . . . . . . . . . . . . . .
Design of Reactive Distillation Processes . .
Procedures for Process Design Studies . . .
Flow sheet for Process Development . . . . .

. 264
.
.
.
.
.
.

265
265
266
267
267
269

Symbols and Abbreviations


aj:
ctj:
E:
fijL:
kj:
kij:
Lj:
Lj1:
NijL:
Nj :
Ntj:
NC:
Q L:
QV:
rij:
SjL:
Vj:
Vj1:
Xi:
xij:
xijI:

specific interfacial area, m /m


mixture molar densities, kmol/m3
energy transfer rate, W
feed flow rate of component i to stage j in
the liquid phase, kmol/s
matrix of mass transfer coefficients, m/s
equilibrium ratio of component i on
stage j
liquid flow rate from stage j, kmol/s
liquid flow rate to stage j, kmol/s
liquid phase mass transfer rate of component i, kmol/s
vector of mass transfer rates, kmol/s
total mass transfer rate, kmol/s
number of compounds
liquid phase heat loss, W
vapor phase heat loss, W
reaction rate of component i on stage j,
kmol/s
ratio of liquid sidewithdrawal
vapor flow rate from stage j, kmol/s
vapor flow rate to stage j, kmol/s
transformed liquid phase composition of
component i
mole fraction of component i in liquid
phase of stage j
liquid mole fraction of component i in
interface

 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim


DOI: 10.1002/14356007.c22_c01.pub2

4.
4.1.
4.2.
4.3.

xj:
Yi:
yij:
yijI:
yj:
DHr:
ni:
n T:

Industrial Applications . . . . . . . . . . . . . . .
Commercial Packing Structures . . . . . . . .
Industrial Catalytic Distillation Processes .
Novel Application of CD with regard to
Process Intensification . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . .

Interface
liquid phase
vapor phase

Subscripts
B:
D:
F:
P:
R:
i:

. 275
. 275

vector of liquid mole fractions


transformed vapor phase composition of
component i
mole fraction of component i in vapor
phase of stage j
vapor mole fraction of component i in
interface
vector of vapor mole fractions
heat of reaction, J/kmol
stoichiometric coefficient for component i
sum of stoichiometric coefficients defined by Equation 9

Superscripts
I:
L:
V:

. 270
. 270
. 272

bottom
distillate
feed
product
reference component
component number

264

j:
t:

Reactive Distillation

stage number
total

Abbreviations and Acronyms


AIBN:
CD:
DIPB:
ETBE:
HB:
HETP:

Azobisisobutyronitrile
Chemical distillation
Diisopropylbenzene
Ethyl tert-butyl ether
High-boiling
Height equivalent to a theoretical plate

1. Introduction
Reactive distillation (RD) is a process in which a
catalytic chemical reaction and distillation (fractionation of reactants and products) occur simultaneously in one single apparatus. Reactive
distillation belongs to the so-called processintensification technologies. From the reaction
engineering view point, the process setup can be
classified as a two-phase countercurrent fixedbed catalytic reactor.
In the literature this integrated reaction
separation technique is also known as catalytic
distillation (CD) or reaction with distillation
(RWD). According to [1], CD is a process in
which a heterogeneous catalyst is localized in a
distinct zone of a distillation column. RD is the
more general term for this operation, which does
not distinguish between homogeneously or heterogeneously catalyzed reactions in distillation
columns. RWD is a trademark of the Koch
Engineering Company for reactive distillation
technology that uses their KataMax packing
structures. A brilliant overview on the current
status of RD technologies, modeling, industrial
applications, etc., can be found in [2].
The present article exclusively deals with RD
processes that operate with a heterogeneous catalyst system, i.e., CD technology. The most
important advantage of CD technology for equilibrium-controlled reactions is the elimination of
equilibrium limitation of conversion by continuous removal of products from the reaction mixture. It is the application of Le Chateliers principle to displace the chemical equilibrium by
increasing the concentrations on the one side of
the reaction, i.e., the reactants, and decreasing it

Vol. 31

LB:
Low-boiling
MB: Medium-boiling
MESH: Material balance/equilibrium condition/
summation equation/heat balance
MTBE: Methyl tert-butyl ether
RD:
Reactive distillation
RWD: Reaction with distillation
TAME: tert-Amyl methyl ether
TAEE: tert-Amyl ethyl ether
TIPB: Triisopropylbenzene

on the other, i.e., the product side. The chemical


composition at this equilibrium point can be
calculated by means of the Gibbs energy of
reaction at a given temperature. Activities must
be used to recalculate the composition from the
equilibrium constant (i.e., the molar fractions of
the components).
Usually, a partially converted reaction mixture, close to chemical equilibrium, leaves the
fixed-bed reactor section and enters the CD
column in the fractionating zone to ensure the
separation of products from feedstock components. The fractionated unconverted feedstock
components enter the catalytic section in the CD
column for additional or total conversion. The
catalyst packing zone is installed in the upper or
lower-middle part of the column, with normal
distillation sections above and below.
CD technology has several advantages over
conventional operating methods, such as a fixedbed reactor connected to a fractionating column,
in which the distillate or bottoms have to be
recycled after further separation steps for a total
overall conversion.
Apart from increased conversion, the following benefits can be obtained [1]:
.

The most important benefit of CD technology


is the lower capital investment, because two
process steps can be combined and carried out
in the same device (so called process intensification). Such integration leads to lower
costs for pumps, piping and electrical
instrumentation.
If CD is applied to an exothermic reaction, the
reaction heat can be used to vaporize part of the
surrounding liquid, which represents three

Vol. 31

fundamental advantages: The maximum temperature in the structured catalytic packing is


limited to the boiling point of the reaction
mixture, so that the danger of hot spots is
reduced significantly (so-called Siedek
uhlung). Also, extremely simple and reliable
temperature control is achieved. In addition,
the integration of reaction heat in the distillation process leads to energy savings by reducing reboiler duty.
Product selectivity can be improved owing to
fast removal of reactants or products from the
reaction zone. Thus, the probability of consecutive reactions, which may occur in the conventional operation mode, is generally
lowered.
If the reaction zone in the CD column is located
above the feed point, poisoning of the catalyst
can be avoided. This leads to longer catalyst
lifetime compared to the conventional mode of
operation.
The possibility to break azeotropes in the vapor
- liquid equilibrium, because reactants or products can act as entrainers or because the
azeotropes can simply disappear.

There are three important constraints for applying CD technology to catalytic chemical
reactions:
.

The use of CD technology is only possible if the


temperature window of the vapor liquid equilibrium is equivalent to the reaction temperature. By changing the column operating pressure, this temperature window can be altered.
The flexibility in the operating temperature of a
CD column is not only restricted by the fact
that two phases are required for the distillation
process. Also the thermal stability of the catalyst can limit the upper operating temperature.
Because of the necessity of wet catalyst pellets,
the chemical reaction must take place entirely
in the liquid phase.
As it is very expensive to change the catalyst in
the structured packing of a CD column, only
catalysts with a long lifetime are suitable for
this process.

As a fourth constraint CD technology is


somehow difficult to model mathematically,
which complicates the scale-up from technical
plant scale to full production scale as well [2].

Reactive Distillation

265

In the literature, it can be found that endothermic reactions are not suitable for the CD technology, because the reaction heat condenses part
of the vapor stream. Although endothermic reactions require more reboiler duty and therefore
exhibit no large energy savings, there are no
restrictions with regard to the application of this
technology [3].
Chemical reactions, which may benefit from
CD technology, should fulfill the above-mentioned criteria in general. Reactions of this type
include, for example, etherifications, esterifications, transesterifications, hydrations, hydrolysis, condensations, hydroisomerizations, oligomerizations, alkylations, transalkylations, and
selective hydrogenations.
An excellent overview of the current status of
published applications is given in [2].

2. Mathematical Modeling of
Reactive Distillation Processes
2.1. Equilibrium-Based Models
Multicomponent separation processes, such as
normal distillation processes, have been modeled
by using the equilibrium-stage concept for a
century. Therefore, early works on reactive distillation also used the equilibrium-stage model to
simulate reactions with superimposed distillation.
The principal assumption of the equilibriumstage model is that the vapor and the liquid stream
that leave the stage are in thermodynamic equilibrium. In most real distillation columns, of
course, the residence time is too short to reach
total equilibrium. For this reason, efficiencies
have been introduced into the model (e.g., Murphree efficiency, vaporization efficiency, etc.) to
account for the nonideal behavior. MESH (i.e.,
Material balance, Equilibrium relationship,
Summation of all substances, and enthalpy balance H) equations are used to simulate conventional distillation columns (! Distillation, 1.
Fundamentals, Section 4.3).
To introduce the chemical reaction superimposed to the distillation further equations are
needed to simulate reactive distillation processes
(! Reaction Columns, Chap. 2.). The simplest
way to consider chemical reactions is to use the
equilibrium constant Ki, but many reactions are
not fast enough to reach chemical equilibrium in

266

Reactive Distillation

Vol. 31

and summation equations (S equations, see


above) for each component i:
X

yIij 1 0;

as well as
X

xIij 1 0

Figure 1. Nonequilibrium based stage model without feed


and side streams

one theoretical stage. Therefore, it is often necessary to use kinetic expressions, which describes
the reaction rate as a function of temperature and
concentration activities of each component of the
reaction scheme.

2.2. Rate-Based Models


To describe the real phenomena of mass transfer
in distillation processes in more detail, a secondgeneration nonequilibrium model was developed
by KRISHNAMURTHY and TAYLOR [4]. A nonequilibrium stage (Fig. 1) represents either a single
tray or a section of packing in a column.
Contrary to the equilibrium-based stage model, this rate-based model treats the vapor and
liquid phases separately and combines them by
means of a mass transfer rate and a heat transfer
rate through the interfacial area. The total mass
transfer rates are:


V
l
V
NjV cV
tj kj aj yj yj Ntj yj



NjL cLtj kjL aj xlj xj NtjL xj

where NjV or NjL are the vectors of molar fluxes of


the vapor or liquid stream, respectively, ctj are the
molar densities, kj are the matrices of mass transfer coefficients, aj is the specific interfacial area,
and NtjV or NtjL are the total mass transfer rate of
the vapor or liquid stream, respectively [5].
At the interface of each stage j phase equilibrium (E equation, see above):
Kij xlij ylij 0

must be fulfilled.
To simulate the reactive distillation process,
an expression for the reaction rate rij f (temperature, concentrations/activities) must be
added. For catalytic distillation the chemical
reaction is taken into account in the material
balance of the bulk liquid phase
1sLj Lj xi;j xi;j1 Lj1 fijL NijL rij 0

where sjL is the liquid side-withdrawal and fijL is


the feed flow rate of component i to stage j in the
liquid phase. For many catalytic distillation processes, liquid catalyst mass transfer must also
be considered.
The reaction heat DHr which is produced
(exothermic reaction) or consumed (endothermic
reaction) is considered in the heat balance of the
liquid phase (see Fig. 1).
YUXIANG AND XIEN [6] used this nonequilibrium stage model with simplified expression of the
reaction kinetics based on concentrations to simulate the synthesis of MTBE, which they investigated experimentally in a column filled with
catalyst granules in cloth pockets. The results of
their simulation agreed fairly well with the experiments. However, SUNDMACHER AND HOFFMANN [7] investigated MTBE formation over a
broad concentration range and found that the
macrokinetics could be correctly described only
by using liquid-phase activities for the intrinsic
reaction kinetics and liquid solid mass
transfer.
The equilibrium-based stage model is an excellent pragmatic approach to simulate reactive
distillation processes, especially those systems
with homogeneous reactions. This model is part
of all available process simulators (e.g., RADFRAC in ASPEN-PLUS, Aspen Technology,
Inc.). For a detailed study of the superimposed
distillation and interactions between reaction with
mass transfer a rate-based approach is usually
preferred (e.g., RATEFRAC in ASPEN-PLUS,

Vol. 31

Reactive Distillation

267

Figure 2. Model complexity in simulation of reactive distillation systems

Aspen Technology, Inc.). Due to its complexity


and the fact that only limited data are available
on mass-transfer coefficients, this approach
does not find a broad usage. Furthermore, small
deviations in these coefficients can have a severe negative impact on the simulation results
[8]. (see Fig. 2)

Xi

Yi

3. Design of Reactive Distillation


Processes

xi  nvi xRR

1 nnT RxR

yi  nvi yRR
1 nnTRyR

where R is a chosen reference component. R can


be reactant or a product component. The stoichiometric coefficient nT in these equations is
defined as:

3.1. Procedures for Process Design


Studies
Only a few systematic methods are available for
the process design of multifunctional units such
as reactive distillation columns. One method is
the concept of reactive distillation lines with
reactive azeotropes by BUZAD AND DOHERTY
[9], BESSLING ET AL. [10] AND STICHLMAIER et al.
[11]. A brief description of this method is given
below. More details can be found in [2].
Figure 3 shows the transformation of coordinates for the equilibrium reaction A B
C.
This transformation reduces the equilibrium concentration of component C to a single point on the
line of the transformed coordinates A and B.
Mathematically one can define a chemical equilibrium reaction by Equation (7):
na Anb B . . .
np P . . . ;

Figure 3. Transformation of coordinates for a equilibrium


reaction A B
C

where ni are the stoichiometric coefficients of the


reactants A, B, etc., and the desired products P,
etc. The transformed liquid and vapor coordinates for an arbitrary component i, i.e., Xi and Yi,
are given by Equation (8) [12].

nT

NC
X

ni

10

i1

Both the vapor and liquid molar fractions of


component P are set to zero, i.e., XP YP 0.
According to Equation (8), the transformed composition coordinates fulfill the unity constraints:
NC
X
i1

Xi

NC
X

Yi 1

11

i1

By using these transformed coordinates for the


feed components A and B, the mass balance and
operating lines of a distillation column can be set
up in the same way as for a system without a
reaction. It reveals that feasible top and bottom
compositions have to meet the lever rule (!
Distillation, 1. Fundamentals, Section 4.1)
through the feed concentration and an appropriate distillation line, as shown in Figure 4. This is
the same procedure as for conventional nonreactive systems.

268

Reactive Distillation

Vol. 31

is increased, a reactive azeotrope appears along


line AB (Fig. 6). According to UNG and DOHERTY
[13] this reactive azeotrope is characterized by
identical vapor and liquid transformed-concentration coordinates (Eq. 12).
Xi Yi

Figure 4. Application of the concept of reactive distillation


lines. As mass balance must be fulfilled, the feed concentration XF is located on a straight line between the transformed
concentrations XD and XB.

Figure 5 shows reactive distillation lines for


the equilibrium reaction A B
C in the
presence of an inert component D. Note that
product C is no longer necessary, because of the
transformed-concentration coordinates between
A and B (see Eq. 8). For a given feed concentration F, two process layouts can be derived. In
design type I, component C decomposes, and B
can be separated from A and D. With design
type II, component C is formed, and the inert
component D is separated from a mixture of A and
B, which is in chemical equilibrium with C. If the
equilibrium constant of the reaction A B
C

12

In this case the reactive azeotrope and the


inert component D are nodes; components A and
B are saddles. Therefore, in contrast to Figure 5,
it is not possible to realize design type I (decomposition of C). Only design type II, i.e., the
formation of C and separation of D at the top
and A, B, and C at the bottom of the RD column,
is feasible. The concept of reactive distillation
lines with reactive azeotropes can also be extended to several other reaction types and to
nonideal reactive distillation systems as well.
In collaboration between the chemical industry and different European universities (EU-research project Brite-Euram), two computational tools have been developed for the design
of RD processes, namely SYNTHESIZER and
DESIGNER. SYNTHESIZER is based on the
above-mentioned analogy between normal distillation processes and RD processes. It investigates the feasibility of RD for an existing process
during re-engineering or in the planning stage for
a new process. DESIGNER gives a more precise

Figure 5. Reactive distillation lines for an equilibrium reaction A B


C in the presence of an inert component D (LB
low-boiling, MB medium-boiling, HB high-boiling)

Vol. 31

Reactive Distillation

269

Figure 6. Reactive distillation lines for an equilibrium reaction A B


C in the presence of an inert component D with a
reactive azeotropic point (LB low-boiling, MB medium-boiling, HB high-boiling).

insight into the RD process with respect to column parameters (e.g., column diameter, etc.). It
contains calculation procedures with equilibrium-based stage models as well as rate-based
models for a specific process design [14].

3.2. Flow sheet for Process


Development
Figure 7 shows a flow sheet as a conceptual basis
for the development of a CD process.
The key feature of this flow sheet is the
feasibility study followed by pilot plant experiments (optional) and a basic engineering
concept. After using the above-mentioned tools
for process synthesis and design (see Section 3.1), the parameters of a kinetic model must
be determined in an appropriate experimental
setup.
Simulation studies with a simple equilibrium-based stage model, such as RADFRAC
(ASPEN-PLUS), should reveal the optimal feed
point to the CD column, the operating conditions of the CD column, and the position of the
reaction zone in the CD column. As a first
approximation, these studies can be carried out
by using equilibrium constants to incorporate
the chemical reaction into the column (less
complex model simulation, see Fig. 2.). For a

more detailed investigation, it is recommended


to use an appropriate activity-based reaction
expression with the determined kinetic parameters in a conventional simulator.
In pilot plant experiments, which are carried
out in a CD column with an inner diameter of
50 100 mm, the simulation studies are examined. As the process is iterative, these experiments are needed to check and to improve the
simulations.
As the last step of the feasibility study, all
obtained results are fed into a final simulation to
decide whether an RD or conventional setup
(e.g., reaction section followed by separation
steps with internal recycle streams) should be
applied. Preferably, a rate-based model, such as
for example RATEFRAC (ASPEN-PLUS), is
used to check whether any mass-transfer limitations interfere with the removal of reactants.
Such limitations can cause undesired consecutive reactions in the reaction zone.
As it is reported by some authors [2] it should
be mentioned that predictive scale-up procedures
from lab-scale experiments (50 80 mm inner
diameter column) to pilot-plant scale (300 mm
inner diameter column) already reveal some
evident deviation in conversion, e.g., for methyl
acetate synthesis or MTBE decomposition in a
CD column. The origin of this effect is up to now
not fully understood.

270

Reactive Distillation

Vol. 31

Figure 7. Flow sheet for the process development of a CD technology (CSTR continuous stirred tank reactor)
(*: see Chapter 1)

The possible reasons for the observed deviations, discussed in [2], could be:
.
.
.

reduced reaction rates due to incomplete catalyst wetting,


mass-transfer limitations, or
maldistribution, etc.

Sensitivity studies of different CD column parameter can help to understand the order of
magnitude of the deviations on the overall con-

version, if scale-up calculations are made. These


studies can facilitate improvement of the reliability of such simulations.

4. Industrial Applications
4.1. Commercial Packing Structures
All commercial CD packing structures have in
common that the catalyst is located in a distinct

Vol. 31

zone of a distillation column. Apart from the


problem of positioning these packing structures,
adequate contact between catalyst surface and
liquid phase must be ensured.
Nowadays, in most industrial processes
macroporous acidic ion-exchange resins are
used, but any other solid catalyst pellet can be
placed in the pockets of a structured packing. As
the entire catalyst structure must be removed
from the distillation column to exchange the
catalyst, one evident prerequisite for the catalyst
is a long lifetime (more than about three years).
Catalyst replacement means shutting down
whole plants which results in a significant loss
of costs and operating time.
CD Tech Catalyst Bales. This type of spatial configuration in a distillation column was
first developed and commercialized by CR & L
(Chemical Research & Licensing Company) in
1980 [15]. It is now licensed by CD Tech (Chemical Distillation Technologies), a joint venture
between CR & L and ABB Lummus Crest. In
this technology the catalyst is supported in a
fiberglass cloth, which is wrapped with stainless
steel wire mesh. The demister wire mesh has two
tasks. First, it stabilizes the fiberglass packing;
second, it provides the void space necessary for
distillation. These fiberglass cloths are rolled into
catalyst bales and stacked on sieve trays in the
column (Fig. 8).

Figure 8. CD Tech catalyst bales

Reactive Distillation

271

KataMax and KataPak Technology.


Analogous catalytic packing configurations were
developed in 1991/1992 independently by Koch
Engineering Company as KataMax technology
[16] and by Gebruder Sulzer [17] as KataPak
technology. Here the catalytic section in the CD
column has a conventional corrugated structure,
wire-cloth distillation packing which contains
appropriate catalyst pellets for the desired reaction. The solid catalyst is held in an envelope of
meshed and crimped screen, sealed at the edges.
These envelopes are stacked together in modular
blocks, so that the crimp of one envelope has a 90
degree angle of inclination to the crimp of the
next. In this architecture, channels are formed for
the rising vapor flow and descending liquid flow.
At the intersections, they provide intensive mixing of vapor and liquid phases and radial distribution in a block. It is reported [18] that the
performance parameters of the KataMax packing
(i.e., the hydraulic and mass transfer efficiencies)
are similar to those of standard distillation devices (e.g., the Flexipac packing). At liquid
loadings of 10 50 m3m2h1 catalyst contact
efficiency exceeded 75 90 %. Catalyst contact
efficiency is the ratio of the rate constant of the
catalyst in the packing to the rate constant of
the catalyst in a stirred tank reactor for a simple
liquid-phase first-order reaction at the same
temperature.
Usually the catalyst content of KataMax or
KataPak lies between 20 and 25 vol. %. For
KataMax a HETP value (height equivalent to a
theoretical plate) between 0.4 and 0.6 m can be
used for simulation. For KataPak the HETP is
1.0 m at normal F factors (! Distillation, 2.
Equipment, Section 2.2). The specific surface
area of KataMax is 210 m2/m3 and of KataPak,
85 or 125 m2/m3 (KataPak-S 170.Y or KataPakS 250.Y, respectively).
The benefits of this type of packing are a very
good distribution of liquid and vapor phases at a
low pressure drop, efficient contact of the reactants
with catalyst pellets and instantaneous distillative
removal of reactants. Therefore, it combines excellent separation performance with an efficient
mass and heat transfer for chemical reaction.
MultiPak Technology. This structured catalyst packing was developed by the company
Montz in 1998 in cooperation with the University
of Dortmund [19]. In MultiPak technology the

272

Reactive Distillation

stacks are built up of alternating layers of


meshed, but not crimped, catalyst envelopes and
crimped structured distillation sheets. From the
viewpoint of reaction and fractionating behavior,
it has the same performance as KataMax or
KataPak. The advantage of MultiPak is the identical geometric structure in an 80-mm laboratory
packing and a 300- or 1000-mm pilot plant
packing. Therefore, scaleup from pilot plant
experiments to industrial plants is much easier.
It is claimed that a defined flow pattern of vapor
and liquid is achieved by this construction. This
has advantages when resins are used as catalyst.
As there is only liquid phase in the catalyst
envelopes, the pellets are completely wetted
under process conditions, and there is no swelling
or shrinkage due to alternating contact with vapor
and liquid. The disadvantage is a higher pressure
drop along the structured packing.
Multichannel Packing [20]. Details on this
type of catalyst packing were published in 2003
by BASF. The point of departure to develop these
packings was to improve trickle-bed reactor performance by installing corrugated distillation
packings inside the reactor and pouring catalyst
particles into them. The catalyst particles are
brought loose into the packing cavities, distributed under the influence of gravity. By alternating packing layers of high and low specific
surface areas (so called catalyst barrier layer)
the catalyst could be held in a specific area of the
CD column. The advantage of these packings is
their simplicity and the ease of catalyst removal
in case of deactivated catalyst. As the way of
introducing the catalyst into a specific column
segment is similar to a normal trickle-bed reactor, the authors describe also a simpler modeling
in case of scale-up studies.
However, if the catalyst size distribution is not
100 % uniform, some particles may enter the
lower part of the CD column. Especially with
reversible reactions, which can only be performed in a specific segment of a CD column
setup (e.g., MTBE synthesis), the undesired reverse reaction could occur in the bottom, if
catalyst particles are present there.
Reactive Rings. The reactive rings of
VEBA Oel [21] are based on packing bodies
such as Raschig rings or Berl saddles. On the
outer and inner surface of these bodies, macro-

Vol. 31

porous ion exchange resins are bonded chemically or physically. For example, the active
Raschig rings are prepared by using an opencelled sintered glass, which is impregnated with
a mixture of styrene, a long chain alkane, pdivinylbenzene, and a radical starter (AIBN:
azobisisobutyronitrile). After the polymerization procedure, the resin is sulfonated with
chlorosulfonic acid to build up the active catalytic sites.
These types of reactive rings have been used
in an etherification process in a pilot plant column [7, 22].
Other Technologies. Another interesting,
but as yet not commercialized, technology for
immobilizing catalyst pellets in a distinct zone of
a CD column is to incorporate a fixed-bed in the
downcomer of a conventional tray column [23].

4.2. Industrial Catalytic Distillation


Processes
Catalytic distillation can be used in catalytic
chemical reactions which are limited by a chemical equilibrium. There are various reactions that
satisfy this criterion, but only for etherification,
esterification, and alkylation (synthesis of ethylbenzene or cumene) is this technology applied on
an industrial scale. An overview of further possible applications can be found in [1, 2, 22].
Etherification Processes (see also !
Methyl Tert-Butyl Ether). The application of CD
technology to etherification is limited to synthesis of methyl tert-butyl ether (MTBE), ethyl tertbutyl ether (ETBE), tert-amyl methyl ether
(TAME), and tert-amyl ethyl ether (TAEE).
Because of their good octane enhancing properties and low volatilities these ethers are excellent gasoline blending compounds (so-called
oxygenates). Reduction of the tetraethyllead content in gasoline in the mid-1970s led to a dramatic
increase in the demand for octane enhancers,
with MTBE being used increasingly. Driven by
the U.S. Reformulated Gasoline Program, which
mandated a minimum oxygen content in gasoline
of 2 %, world MTBE production has reached ca.
19  106 t/a in 1997 [24]. The successful use of
CD technology in MTBE production led to its
worldwide establishment.

Vol. 31

Reactive Distillation

273

Table 1. Overview of CD processes for ether production [27]

Process (licensor)

Process description

Catalyst

ETHERMAX (UOP)

fixed-bed tubular reactor or adiabatic recycle reactor


followed by a reactive distillation column with KataMax Packing
adiabatic reactor operating at the boiling point
followed by a reactive distillation column with CD Tech Bales
adiabatic fixed-bed reactor operating at the boiling
point followed by a reactive distillation column with CD Tech Bales
recycle reactor, fixed-bed reactor followed by a
reactive distillation column

sulfonated
ion-exchange resin
acidic
ion-exchange resin
trifunctional
ion-exchange resin
acidic ion-exchange resin

CDMTBE & CDTAME


(CD-Tech)
CDETHEROL
(CD-Tech)
CATACOL (IFP)
*

Market
share*
ca. 32 %
ca. 58 %

ca. 10 %

For MTBE, produced by CD technology in 1997

It should be mentioned that end of 1997


California started to ban MTBE from the gasoline pool due to its effect as a water contaminant
and its possible cause of health problems (see
also ! Methyl Tert-Butyl Ether). In mid-2006
the U.S. Environmental Protection Agency terminated the legal requirement of using MTBE in
the gasoline pool [25]. The production capacities
of MTBE are therefore shrinking day-by-day.
Table 1 summarizes processes for ether production based on CD technology. As shown in the
simplified process flow diagram for the ETHERMAX process (Fig. 9), all processes have in
common a reactor section followed by a CD
column. There is no industrial process in which
the whole conversion is performed in a single CD
column. A partially converted mixture from the
reactor section, which is nearly in chemical

equilibrium, enters the CD column below the


catalyst packing zone to ensure the separation of
the ether from the feed stream. The catalyst
packing is installed in the upper middle of the
column, with normal distillation sections above
and below.
Usually, the reaction of the isoalkene with
methanol or ethanol is conducted in the presence
of a slight stoichiometric excess of the alcohol. In
addition to the advantages of a higher selectivity
for the ether (lower formation of C8 and C10
dimers) and shifting of the chemical equilibrium
to the product side, this leads to a secure process
control. In the absence of the alcohol in the
reaction zone of the CD column exothermic
dimerization and oligomerization of the C4 and
C5 alkenes takes place at high reaction rates. A
sharp and excessive temperature rise (hot spot)

Figure 9. Simplified flow diagram for the ETHERMAX process

274

Reactive Distillation

causes irreversible catalyst deactivation and catalyst damage. The excess alcohol is collected in
the overhead of the CD column and separated
from the hydrocarbon stream by extraction with
water.
In the case of MTBE production, a conversion
of up to 99.9 % is possible by using this technology, if sufficient ether is present at the bottom of
the column. Because of a minimum boiling azeotrope in the binary MTBE methanol system,
the ether carries the alcohol into the reaction zone
of the CD column. This unusual vapor liquid
behavior and the high difference in the heat of
vaporization between C4 hydrocarbons and
methanol are possibly responsible for the known
multiple steady states in CD column operation
[22, 26].
The TAME process usually gives a lower
conversion of 91 95 % only.
Esterification Processes. Esterification is a
good example for the beneficial use of the CD
technology. In conventional methyl acetate production the recovery of methyl acetate from the
reactor outlet stream is complicated, because
methanol forms a low boiling azeotrope with
water and the separation of water from unconverted acetic acid is difficult. Due to these separation problems, the old Eastman Kodak process
used a reactor coupled with eight distillation
columns and an extraction column [28].
In the reactive distillation process, almost
pure methyl acetate can be collected in the
overhead of a single CD column at acetic acid

Figure 10. Methyl acetate synthesis with CD technology

Vol. 31

conversions of greater than 99 %. The Eastman


Kodak process uses sulfuric acid as catalyst [29],
but a heterogeneous catalyst system (acid ionexchange resin) can also be used successfully
[30].
The conceptual basis for the successful implementation of reactive distillation in methyl
acetate synthesis is shown in Figure 10. There
are four zones in the column, which ensure the
fairly high conversion. Acetic acid is separated
from methyl acetate at the top of the column
(zone I). In zone II, an extractive distillation
section below the acetic acid feed extracts water
from methyl acetate. The reaction takes place in
the middle of the column (zone III). At the
bottom (zone IV) methanol is fed and stripped
from descending byproduct water.
Alkylation Processes. CD-technology is also applied for alkylation reactions in the case of
ethylbenzene and cumene manufacture [1]. As
the two processes are fairly similar, only the
cumene process is discussed here.
In the CDCUMENE process, CD Tech catalyst bales are filled with an acidic, wide-pore
zeolite catalyst, for example, zeolite Y, b or w
[31]. The bales are stacked in the middle of a
distillation column. Below the reaction zone
propene is fed; benzene is passed as reflux to
the top of the column. The propene concentration
in the reaction zone is held fairly low to slow
down a side reaction in which diisopropylbenzene (DIPB) and triisopropylbenzene (TIPB) are
formed. The cumene product with impurities of

Vol. 31

Figure 11. Transesterification of methyl acetate to butyl


acetate with CD technology in a divided wall column

ethylbenzene and n-propylbenzene, is taken from


the bottom of the column. The overhead pressure
of the CD column is maintained at 5 bar.
The advantages of CD technology in cumene
synthesis are lower formation of oligo-isopropylbenzenes (DIPB and TIPB), higher catalyst
lifetime, and a higher conversion of benzene
[32]. The internal benzene recycle to the reaction
zone, which is a result of the distillative cumene benzene separation in the lower portion
of the column, ensures the observed higher
conversion.

4.3. Novel Application of CD with


regard to Process Intensification
Distillation columns with dividing walls are used
with success in some chemical processes. An
EU-research project named INSERT has been
started to investigate if CD technology can also
be applied advantageously to such distillation
columns with divided wall internals [33].
The successful use of CD technology in a
divided wall column is reported in [34] for the
transesterification of methyl acetate to butyl
acetate (Fig. 11.). This process can be performed
in one column of this type instead of using one
CD column followed by two normal distillation
columns.

References
1 G. G. Podrebarac, F. T. T. Ng, G. L. Rempel, CHEMTECH 1997, 37 45.

Reactive Distillation

275

2 K. Sundmacher, A. Kienle (eds.): Reactive Distillation,


Wiley-VCH, Weinheim 2003.
3 Huls, EP 0 726 241, 1995 (M. Sakuth, U. Peters).
4 R. Krishnamurthy, R. Taylor, AIChE. J. 31 (1985) 449
456.
5 R. Taylor, R. Krishna: Multicomponent Mass Transfer,
Wiley, New York 1993.
6 Z. Yuxiang, X. Xien, Trans. IchemE 70 (1992) 465
470.
7 K. Sundmacher, U. Hoffmann, Chem. Eng. Sci. 49 (1994)
3077 3089.
8 A. P. Higler, R. Taylor, R. Krishna, Chem. Ing. Sci. 54
(1999) 1389 1395.
9 G. Buzard, M. F. Doherty, Computers Chem. Engng. 19
(1995) 395 408.
10 B. Bessling, G. Schembecker, K. H. Simmrock, Ind. Eng.
Chem. Res. 36 (1997) 3032 3042.
11 J. Stichlmaier, T. Frey, Chem. Ing. Tech. 70 (1998)
1094 1095.
12 J. Espinosa, P. A. Aguirre, G. A. Perez, Ind. Eng. Chem.
Res. 34 (1995) 853 861.
13 S. Ung, M. F. Doherty, AIChE J. 41 (1995) 2383 2392.
14 G. Schembecker et al., Chem. Ing. Tech. 70 (1998) 1096.
15 CR & L, US 4 232 177, 1980 (L. A. Smith Jr.
16 Koch Engineering Company, EP 0 428 265, 1991 (A. P.
Gelbein, M. Buchholz).
17 Gebruder Sulzer, EP 0 396 650, 1992 (R. Shelden, J.-P.
Stringaro).
18 J. L. DeGarmo, V. N. Parulekar, V. Pinjala, Chem. Eng.
Prog. (1992), no. 3, 43 50.
19 Gebrauchsmuster No. 298 7 007.3, 1998 (A. Gorak, L.
U. Kreul).
20 BASF, WO 03/047747 A1, 2003 (G. Kaibel et al.).
21 VEBA Oel, US 5 244 929, 1993 (K. Gottlieb, W. Graf, K.
Schaedlich, U. Hoffmann, A. Rehfinger, J. Flato).
22 K. Sundmacher, Dissertation, Universitat Clausthal-Zellerfeld, 1995.
23 Koch Engineering Company, EP 0 664 721, 1997 (N.
Yeoman, R. Pinaire, M. A. Ulowetz, T. P. Nace, D. A.
Furse).
24 Chem. Week, Oct. 29, 1997, 44.
25 ICIS Chemical Business 2006, Mar./Apr., 32.
26 S. Hauan, T. Hertzberg, K. Lien, Ind. Eng. Chem. Res. 34
(1995) 987 991.
27 Hydrocarbon Processing 73 (1992) 104 110.
28 B. Bessling, Dissertation, Universitat Dortmund, 1998.
29 V. H. Agreda, L. R. Partin, W. H. Heise, Chem. Eng. Prog.
(1990) no. 2, 40 46.
30 J. Gmehling, J. Krafczyk, Chem. Ing. Tech. 66 (1994)
1372 1375.
31 G. R. Meima, M. J. M. van der Aalst, M. S. U. Samson, J.
M. Garces, J. G. Lee, Catalyst of Acid and Bases, Proc.
DGMK-Conf. (1996) 125 137.
32 CR & L, US 4 849 569, 1989 (L. A. Smith Jr.
33 E. Geiler, C. Gromann, S. Sander, C. Flisch, O. Ryll, H.
Hasse, Chem. Ing. Tech. 78 (2006) 1282 1283.
34 G. Kaibel, C. Miller, T. Holtmann, H. Schoenmakers,
Chem. Ing. Tech. 77 (2005) 1749 1758.

276

Reactive Distillation

Further Reading
C. A. M. Afonso: Green Separation Processes, Wiley-VCH,
Weinheim 2005.
Z. Lei B. Chen Z. Ding: Special Distillation Processes, 1st
ed., Elsevier, Amsterdam 2005.

Vol. 31
W. L. Luyben, C.-C. Yu: Reactive Distillation Design and
Control, Wiley, Hoboken, NJ 2008.
K. Sundmacher: Reactive Distillation, Wiley-VCH, Weinheim 2003.

You might also like