Anda di halaman 1dari 14

Ind. Eng. Chem. Res.

2006, 45, 6007-6020

6007

SEPARATIONS
A Method for Modeling Two- and Three-Phase Reactive Distillation Columns
Rahman Khaledi and P. R. Bishnoi*
Department of Chemical and Petroleum Engineering, The UniVersity of Calgary, 2500 UniVersity DriVe NW,
Calgary, Alberta, Canada T2N 1N4

An algorithm for the steady-state simulation of two- and three-phase multistage reactive distillation processes
with equilibrium chemical reactions is developed. In the developed algorithm, the phase stability, phase
equilibrium, and chemical reaction equilibrium calculations are preformed simultaneously. The algorithm
was used to simulate a variety of two- and three-phase reactive distillation processes. The simulation results
were compared against the available experimental data in the literature. Good agreement was observed between
the simulation results and experimental data. A multistage reactive distillation process for the production of
cyclohexanol from the hydration of cyclohexene with two liquid phases present on multiple stages was
simulated. A high-purity cyclohexanol product with a high cyclohexene conversion was obtained for this
process.
Introduction
Reactive distillation is a process in which the chemical
reaction and separation happen continuously in a single operation unit. It is generally used for performing equilibrium-limited
chemical reactions. This technology offers significant benefits
over conventional processes, such as the elimination of a
separate reaction vessel, fewer separation units, high conversion
of reactants, improved selectivity of products, and reduced
reboiler duty in the case of exothermic reactions.
The reactive distillation process has been long known in the
chemical industry.1 However, it is only during the past decade
that there has been a tremendous amount of interest and an
increase in the number of publications on this subject.2 This
process recently became very important in the production of
fuel additives such as methyl tert-butyl ether (MTBE), ethyl
tert-butyl ether (ETBE), and tert-amyl methyl ether (TAME),
and also in the production of many other chemicals such as
esters and alcohols.3 Most of the published papers on the
modeling of reactive distillation primarily involve the development of methods for two-phase (vapor-liquid) reactive distillation columns. Taylor and Krishna2 have provided an excellent
review on the recent developments in the modeling of reactive
distillation processes.
Even though there has been a considerable amount of research
on the modeling of reactive distillation systems, there has been
a little research done on reactive distillation processes involving
liquid-phase splitting. Liquid-phase splitting may occur on one
or more stages within the reactive distillation column when
highly nonideal mixtures are being processed. An example of
such cases is the heterogeneous azeotrope reactive systems. In
this case, a low-boiling azeotrope vapor mixture is formed in
the top section of the reactive distillation column. This vapor
mixture splits into two liquid phases in the decanter after being
condensed in the condenser. Hexyl acetate4 and butyl acetate5-8
production via reactive distillation are two examples of the
* To whom correspondence should be addressed. Tel: +(403) 2206695. Fax: +(403) 282-3945. E-mail address: bishnoi@ucalgary.ca.

heterogeneous azeotrope reactive distillation process. In other


cases, the presence of partially miscible components in the
system results in liquid-phase splitting under certain conditions.
In such systems, two liquid phases might be found on several
stages, depending on the operating conditions of the column.
An example of such cases is the production of cyclohexanol
via a reactive distillation process.9,10 In this system, the presence
of organic compounds (cyclohexanol, cyclohexene, and cyclohexane) that are partially miscible in water cause the liquid to
split into aqueous and organic liquid phases.
In recent studies on the simulation of butyl acetate reactive
distillation columns, to model the liquid-phase splitting in the
column, only a liquid-liquid phase equilibrium model was
considered for the decanter.5,8 Schmitt et al.4 conducted an
experimental study for the production of hexyl acetate via
reactive distillation. In their simulation of experiments, they did
not model the decanter, and the reflux to the column was
considered to be an independent feed with the conditions similar
to the measured data from the experiments. Qi and Sundmacher11 and Qi et al.9 developed a systematic approach to study
the feasibility of reactive distillation for the systems with liquidphase splitting using residue curve maps. The residue curve
maps analysis of Steyer et al.,10 from the same research group,
on a batch reactive distillation for the production of cyclohexanol, shows that the synthesis of cyclohexanol via a continuous
multistage three-phase reactive distillation column is conceptually feasible. Gumus and Ciric12 proposed a bilevel optimization
approach to explore the design of reactive distillation columns
that may involve liquid-phase splitting on the stages. Their
proposed approach includes a Gibbs free energy minimization
problem that determines the number of phases present on the
stages. This Gibbs free energy minimization is nested in a larger
minimization problem that minimizes the annual cost of the
column. Gumus and Ciric12 demonstrated the computational
performance of their algorithm by solving a single-stage
multiphase reactive flash example.
The modeling of multistage three-phase (VLL) reactive
distillation columns involves many complexities, because of the

10.1021/ie051384a CCC: $33.50 2006 American Chemical Society


Published on Web 07/26/2006

6008

Ind. Eng. Chem. Res., Vol. 45, No. 17, 2006

Figure 2. Schematic of a reactive stage in multiphase reactive distillation


column (k ) 1 denotes vapor; k ) 2, ..., denotes liquid).

developed by Gupta et al.,14 in the governing equations of the


system. The developed algorithm15 is used successfully to
simulate a variety of two- and three-phase reactive distillation
columns.
Model Equations for Multiphase Reactive Distillation
Columns

Figure 1. Schematic representing a multistage three-phase reactive


distillation column.

simultaneous presence of reaction, liquid-phase splitting, and


separation in the system. The main difficulty in the modeling
of these processes is due to the fact that the number of phases
present on each stage is not known in advance. To our
knowledge, there are no published articles in the open literature
on the simulation of three-phase multistage reactive distillation
columns.
In the present work, the development of a method for the
simulation of two- and three-phase multistage reactive distillation columns with chemical reaction equilibrium on reactive
stages is presented. This method is an extension of the threephase distillation without reaction developed recently by Khaledi
and Bishnoi.13 In the developed method, the liquid-phase
splitting is allowed at each stage and a prior knowledge of the
phase pattern is not required. The phase stability test, and the
phase and chemical reaction equilibrium calculations, are
conducted simultaneously. The simultaneous phase stability test
is accomplished by including a phase stability equation,

The schematic of a multistage three-phase reactive distillation


column is illustrated in Figure 1. The column consists of Ns
theoretical stages. The stages are numbered from the top to the
bottom of the column. The first theoretical stage is combination
of a condenser and a decanter, and the last theoretical stage is
a partial reboiler. The condenser can be either a partial or total
condenser. Figure 2 shows the schematic of a typical equilibrium
reactive stage for the column. Multiple liquid phases may exist
on the stage. An independent feed stream consisting of Nc
components is introduced on the stage and Nr number of
chemical reactions occur on the stage. The vapor and liquid
side streams may be withdrawn from the stage and heat may
be withdrawn or added to the stage, using a side heater or cooler.
The following assumptions were made in formulating the
mathematical model equations:
(1) All the phases are completely separated from each other
and are at thermodynamic equilibrium when leaving the stage.
(2) Chemical reaction equilibrium conditions exist on the
reactive stages.
(3) The process occurs under steady-state conditions.
The governing equations that are required to describe a
reactive distillation column are component material balance,
chemical reaction equilibrium, energy balance, mole fraction
summation, phase fraction summation, phase stability, and
thermodynamic phase equilibrium. These equations are written
for steady-state conditions in the form of discrepancy equations
for each stage as follows. It is noted that the model equations
for reactive distillation are similar to distillation without
reaction,13 except for the component material balance (eq 1)
and chemical reaction equilibrium (eq 2).
Component Material Balance Equation. The component
material balance around stage j for component i can be written
as

M
ij ) Fjzij -

nijk + (1 - wrj+1,1)ni,j+1,1 +

k)1

k)2

Nr

irjr

r)1

(1 - wrj-1,k)ni,j-1,k + j

(i ) 1, ..., Nc; j ) 1, ..., Ns) (1)


The last term in eq 1 corresponds to the total number of moles

Ind. Eng. Chem. Res., Vol. 45, No. 17, 2006 6009

of component i generated or consumed by the chemical reaction.


i,r is the stoichiometric coefficient of component i in reaction
r and j,r is the extent of reaction r on stage j. The stoichiometric
coefficient is negative for the reactants and positive for the
products. The parameter j is set to 1 for a reactive stage and 0
for a nonreactive stage.
Chemical Reaction Equilibrium Equation. In the model,
it is considered that Nr number of reactions are occurring in the
system. It is noted that, at thermodynamic equilibrium, the
activity of a component is equal in all liquid phases, provided
that the same standard state fugacity of the component is used
for all the liquid phases. Under this condition, the equation for
the chemical reaction equilibrium is the same for all the liquid
phases. Therefore, the chemical reaction equilibrium equation
written for only one liquid phase is enough to describe the
chemical reaction equilibrium condition on a reactive stage. The
chemical reaction equilibrium for liquid-phase k ) 2, based on
liquid-phase activity for the reactive stage j, is given by

Rrj ) (

i a ijk| |)reactants ir

1
Keq
rj

i a ijk| |)products

St
jk )

Rjkjk
Rjk + jk

(j ) 1, ..., Ns; k ) 2, ..., )

where Krjeq is the chemical equilibrium constant for reaction r


on stage j. The chemical reaction equilibrium equation is valid
only for the reactive stages where the chemical reaction occurs.
Hence, the total number of equations describing chemical
reaction equilibrium is the product of the number of reactions
and the number of reactive stages (i.e., Nrs Nr).
Energy Balance Equation. In formulating the energy balance
equation, the enthalpies are calculated using the reference state
of pure elements, as discussed later. As a result, the heat of
reaction is embedded in the enthalpy of the component and does
not appear as an explicit term in the energy balance equation.
The energy balance equation for stage j is given by

Note that the reference phase (k ) 1, which denotes the vapor


phase) is always stable and the phase stability variable, jk, for
this phase is always equal to zero. Hence, the phase stability
equation for this phase is excluded from the model equations.
Phase Equilibrium Equation. The thermodynamic phase
equilibrium equation is given by

xijk ) xij1Kijkejk

(i ) 1, ..., Nc; j ) 1, ..., Ns; k ) 2, ..., )


(7)

where Kijk is the equilibrium ratio, which is defined as the ratio


of fugacity coefficient of the reference phase and the fugacity
coefficient of phase k. That is,

Kijk )

ij1
ijk

(i ) 1, ..., Nc; j ) 1, ..., Ns; k ) 2, ..., )

nijk ) ijkmij

mij )

(j ) 1, ..., Ns) (3)

Mole Fraction Summation Equation. The mole fraction


summation for phase k on stage j, in terms of discrepancy, is
given by

(10)

nijk
(11)

nijk

ni,j+1,1hhi,j+1,1 +

nijkhhijk + (1 - wrj+1,1)
i)1 k)1
i)1

(1 - wrj-1,k)ni,j-1,khhi,j-1,k
i)1 k)2

nijk

k)1

Nc

Nc

(9)

where mij and ijk are given by

)
Nc

(8)

The component molar flow rate, nijk, may be expressed in


terms of the total number of moles of component i leaving stage
j (mij) and variable ijk as follows:

ijk )

Hfj + Qj -

(6)

ir

(r ) 1, ..., Nr; j ) 1, ..., Ns; k ) 2) (2)

Ej

j in the form of a discrepancy equation:

k)1

By combination and manipulation of eqs 7 and 11, an expression


for variable ijk is obtained as follows:

ijk )

RjkejkKijk
(12)

Rjpe Kijp

p)1
jp

Nc

Sjk )

(xijk - xij1)

i)1

(j ) 1, ..., Ns; k ) 2, ..., ) (4)

where xij1 refers to the mole fraction in the reference phase (k


) 1, which denotes the vapor phase).
Phase Fraction Summation Equation. The phase fraction
summation equation is given as

PS
j )

Rjk - 1

(j ) 1, ..., Ns)

(5)

k)1

Phase Stability Equation. The modified form of the phase


stability equation suggested by Gupta et al.14 is written for stage

The component mole fraction, xijk, can be expressed as

xijk )

mijKijkejk
Nc

(13)

mljRjpe Kijp

l)1
p)1
jp

Derivation of eqs 12 and 13 is given by Khaledi and Bishnoi.13


Equations 12 and 13 implicitly incorporate the equilibrium
relation given by eq 7. Equation 13 is used to update the phase
composition in the outer loop. The governing equations can be
expressed in terms of mij and ijk. This simplifies the form of
governing equations and their derivatives. The new set of

6010

Ind. Eng. Chem. Res., Vol. 45, No. 17, 2006

governing equations in terms of variables mij and ijk is given


below:
Component Material Balance Equation:

M
ij ) -mij + Fij + (1 - wrj+1,1)i,j+1,1mij+1 +
Nr

irjr

r)1

(1 - wrj-1,k)i,j-1,kmij-1 + j

k)2

(i ) 1, ..., Nc; j ) 1, ..., Ns) (14)


Chemical Reaction Equilibrium Equation:

Rrj ) (

i a ijk| |)reactants ir

Keq
rj

i a ijk| |)products
ir

(r ) 1, ..., Nr; j ) 1, ..., Ns; k ) 2) (15)


Energy Balance Equation:
Nc

Ej ) Hfj + Qj -

ijkmijhhijk +
i)1 k)1
Nc

(1 - wrj+1,1)

i,j+1,1mi,j+1hhij+11 +

i)1

Nc

(1 - wrj-1,k)i,j-1,kmi,j-1hhij-1,k
i)1 k)2

(j ) 1, ..., Ns) (16)

Mole Fraction Summation Equation:

Sjk )

1
Nc

l)1

Nc

ijk

i)1 R

mlj

jk

Figure 3. Flowchart of the inside-out algorithm.

ij1
Rj1

mij

(j ) 1, ..., Ns; k ) 2, ..., )

(17)

Model Equations Solution Algorithm

Phase Fraction Summation Equation:

PS
j )

Rjk - 1
k)1

(j ) 1, ..., Ns)

(18)

Phase Stability Equation:

St
jk )

Rjkjk
Rjk + jk

(j ) 1, ..., Ns; k ) 2, ..., )

(Tj), phase fractions (Rjk), and phase stability variables (jk).


This results in Ns (Nc + 2) + Nrs Nr number of unknown
variables.

(19)

Equations 14-19 are model equations for the steady-state


simulation of multistage, multicomponent, and multiphase
reactive distillation columns. The variable ijk in the aforementioned model equations is calculated using the phase equilibrium
relation given by eq 12. These model equations describe twoand three-phase reactive distillation columns under chemical
reaction equilibrium conditions. The advantage of the aforementioned set of equations is that they easily allow for the
appearance and disappearance of liquid phases during computations.13 The total number of these equations is Ns (Nc + 2)
+ Nrs Nr. For the specified feed conditions, side heat duties,
side-withdrawal ratios, and stage pressures, we are required to
solve the governing equations for the overall component stage
flow rates (mij), stage reaction extents (j,r), stage temperatures

An approach that is similar to the inside-out method of


Boston and Sullivan16 is implemented for solution of the model
equations.13 Figure 3 shows the flow diagram of the solution
algorithm. In this approach, the thermodynamic properties are
approximated with simple models. These models are used to
evaluate the thermodynamic properties in an inner loop. When
the inner loop is converged, the parameters of approximate
models are updated using rigorous thermodynamic models in
an outer loop. The component activity, using a liquid activity
model, is expressed as

a ik ) ikxik

(20)

In the aforementioned equation, the activity coefficient is


assumed to be only a function of temperature in the approximate
thermodynamic model. This temperature dependency is given
by

ln ik(T) ) ika - ikb

1
1
T Tb

(21)

The approximate model for the K-values and the partial molar

Ind. Eng. Chem. Res., Vol. 45, No. 17, 2006 6011

residual enthalpies are presented as simple temperature-dependent functions in the following forms:

ln Kik(T) ) Kika - Kikb

1
1
T Tb

(22)

hhRik ) hika + hikb(T - Tb)

(23)

It is assumed that the parameters of the approximate models


ika, ikb, Kika, Kikb, hika, and hikbsare independent of temperature
and composition and are kept constant in the inner loop. These
parameters are evaluated from rigorous thermodynamic models
in the outer loop. The material balance, chemical reaction
equilibrium, energy balance, mole fraction summation, phase
fraction summation, and stability equations (eqs 14-19) form
the set of inner loop equations. These equations are grouped
by the stage,17 as given by eqs 24 and 25:

F ) {f1, f2, ..., fj, ..., fNs}T

(24)

partial molar residual enthalpy of component i. The standard


heat of formation, hi,f298, is calculated for the formation of a
compound from its pure elements, which are chosen as a
reference state. The enthalpy of a pure component at the ideal
gas state and temperature T, relative to the reference temperature
(To) of 298 K, is calculated using the ideal-gas heat-capacity
data. That is,

hig
i )

S
S
St
St
St
, ..., j,
, PS
j,k
j , j,2, ..., j,k, ..., j,} (25)

The independent variable vector, which consists of overall


component molar flow rates, extent of reactions, inverse of
temperatures, phase stability variables, and phase fractions, is
given by

X ) {x1, x2, ..., xj, ..., xNs}T

(26)

where xj is the independent variable vector for stage j, in which


the variables are sorted in the following order:

1
xj ) m1,j, m2,j, ..., mNc,j, j,1, j,2, ..., j,Nr, ,
Tj

hhRi ) -RT2

Keq
r )

The set of Ns (Nc + 2) + Nrs Nr inner loop equations,


F, are solved simultaneously for Ns (Nc + 2) + Nrs Nr
independent variables, X, using a combination of homotopycontinuation18 and Powells method.19 The phase compositions
are updated using eq 13 in the outer loop. Consequently, the
parameters of approximate model are updated using the
calculated compositions. A more-detailed description of the
method of solution of the model equations and the convergence
criterion are given by Khaledi and Bishnoi.13

Keq
r ) a exp

hRi
hhi ) hi,f298 + hig
i +h

(28)

In eq 28, hhi is the partial molar enthalpy, hi,f298 the standard


hRi the
heat of formation, hig
i the ideal gas-state enthalpy, and h

p,x

(30)

ir

(27)

The thermodynamic properties are calculated using an equation of state for the vapor phase and a liquid activity model for
the liquid phase.
Enthalpy Calculations. The enthalpy of a stream is calculated using the partial molar enthalpies of its components. The
partial molar enthalpy of a component i is calculated using the
following equation:

ln i
T

i a ijk| |)products

Thermodynamic Properties Calculations

( )

Chemical Equilibrium Constant Calculations. The chemical equilibrium constant can be either evaluated from standardstate Gibbs free-energy data or from chemical equilibrium
experimental measurements. In the latter method, the liquid
compositions are measured at the chemical equilibrium condition
for a specific temperature. Using an appropriate liquid activity
model, the activities of the components are evaluated in the
reaction mixture, and the chemical equilibrium constant is
calculated from eq 31. The temperature dependency of the
equilibrium constant can be expressed by eqs 32 and 33. The
parameters a, b, c, d, e, and f in these equations are determined
by fitting the equation to the experimental equilibrium constant
data obtained at different temperatures:

j,2, ..., j,k, ..., j,, Rj,1, ..., Rj,k, ..., Rj,

(29)

The equations for the calculation of the component partial molar


residual enthalpy hhRi in the vapor and liquid streams are
derived from the expressions for the fugacity coefficients, using
an equation of state for the vapor phase and a liquid activity
model for the liquid phase. The following equation is used for
the derivation of the component partial molar residual enthalpy,
hhRi :

where fj is the discrepancy of eqs 14-19 for stage j, which are


sorted in the following order:
M
M
R
R
S
fj ) {1,j
, 2,j
, ..., NMc,j, 1,j
, 2,j
, ..., NRr,j, Ej , j,2
, ...,

TT)298K Cpigi dT

(31)
a ijk|ir|)reactants

(Tb + c ln T + q(T))

(32)

where

q(T) ) dT + eT2 + fT3

(33)

In the Simulation Results and Discussions section, which is


presented later in this paper, the expression and the parameters
for chemical equilibrium constant Keq
r , are given for each of
the reactions involved in the reactive distillation examples.
Initialization of the Iterative Variables. To begin the
iterative calculations for solving the model equations, initial
values for all of the iterative variables are required. The
initialization procedure for two-phase reactive distillation
columns begins by initializing its temperature profile. This
profile is assumed to be a linear interpolation between the
condenser bubble point and reboiler dew point temperatures
evaluated at the average composition of all feeds introduced to
the column. A linear pressure profile is considered in the
column. Vapor and liquid-phase fractions are initialized by
assuming constant molal overflow in the column and initial
phase-stability variables for all phases are set equal to zero. To
evaluate the initial values for the component overall flow rates,

6012

Ind. Eng. Chem. Res., Vol. 45, No. 17, 2006

mij, first a nonreactive form of eq 14 (by dropping the reactive


term from the equation) is utilized. The component overall flow
rates mij are calculated by solving the tridiagonal system of linear
equations formed by this nonreactive form of component
material balance equation. The component phase fractions, ijk,
in eq 14 are calculated using eq 12 with ideal equilibrium ratios
evaluated at stage temperatures and pressures. Subsequently,
the component molar flow rate, nijk, is calculated using eq 9.
To estimate the initial number of moles of the components in
the presence of chemical reaction equilibrium on each stage,
the component material balance and chemical reaction equilibrium equations, given by eqs 34 and 35 are solved simultaneously for each reactive stage.
Nr

ni ) ni +

i,rr

r)1

(i ) 1, ..., Nc)

(( )) (( ))
i

ni

|ir|

Nc

nl

l)1

i
eq

Kr

reactants

ni

|ir|

Nc

nl

l)1

(34)

)0

products

(r ) 1, ..., Nr) (35)

In eq 34, the initial number of moles, ni, are assumed to be the


liquid component molar flow rates in the liquid phase leaving
stage j that were calculated in the previous step. Solution of
the system of nonlinear equations (eqs 34 and 35) using the
Newton-Raphson method for each stage gives the number of
moles of each component, ni, in the liquid phase and the extent
of reactions, r, for each reactive stage. Note that eq 35 is a
simplified form of the chemical reaction equilibrium equation
(eq 15) obtained by assuming ideal solution and replacing
activities of the components by mole fractions (i.e., assuming
an activity coefficient equal to 1). This simplification is
acceptable because the obtained extent of reactions and component molar flow rates are used only as initial estimates for
these variables. Subsequently, the initial values for the component overall flow rates mij are recalculated using eq 10. The
initial liquid-phase component mole fractions are calculated from
liquid component mole numbers, and the vapor-phase component mole fractions are simply calculated using thermodynamic
equilibrium relation with ideal K-values. The parameters for
the approximate thermodynamic models then are calculated at
the initial stage temperatures and compositions, using the
rigorous thermodynamic models.
In a three-phase reactive distillation column, iterative calculations require a good set of initial estimates. To obtain these
initial estimates, the column first is initialized as a two-phase
reactive distillation column. By performing one iteration of twophase reactive distillation calculations, appropriate initial temperature, composition, and flow rate profiles are obtained. To
generate the composition and phase fraction for the second liquid
phase on the stages, a liquid-liquid flash calculation is
performed on the liquid phase that was obtained from the twophase reactive distillation calculations in the previous step. The
initial K-values for the liquid-liquid flash calculation are
generated using a noniterative stability analysis method.20 The
resulting compositions and phase fractions from the liquidliquid flash calculation are used as initial values for the liquid
phases. If the performed liquid-liquid flash calculation does
not find a second liquid on a stage, the compositions and the
phase fractions of a nearest stage that has two liquid phases are
used to initialize the compositions and phase fractions of the
liquid phases on this stage. These compositions are used to

initialize the parameters of the approximate thermodynamic


models using the rigorous thermodynamic models for the first
outer-loop iteration in the three-phase reactive distillation
column.
Simulation Results and Discussions
To demonstrate the capability of the developed algorithm in
modeling a variety of reactive distillation processes, the simulation results of five different reactive distillation columns
examples are presented in this article. Table 1 gives the details
of column specifications for the reactive distillation examples.
The first and second examples are two-phase reactive distillation
columns for the production of methyl acetate and ethyl tertbutyl ether (ETBE), respectively. The third example is a
heterogeneous azeotropic reactive distillation column for the
production of butyl acetate. The liquid-phase splitting occurs
in the decanter of this column, which makes the process a threephase reactive distillation type. To examine the validity of the
simulation results, the results of the first and third examples
are compared with the available experimental measurements
obtained from the literature. Example 4 is another heterogeneous
azeotropic reactive distillation column that is used for the
production of hexyl acetate. The last example is a three-phase
reactive distillation column for the production of cyclohexanol.
The presence of partially miscible components in the system
results in the formation of two immiscible liquid phases on many
of the column stages. To our knowledge, this is the first time
that a multistage reactive distillation column with two liquid
phases present on the stages within the column is simulated.
For all the examples discussed in this work, the PengRobinson21 equation of state is used to calculate the vapor phase
thermodynamic properties. For the liquid phase, liquid activity
models are used as mentioned in the discussions for each
example. The ideal-gas heat capacities and standard-state
fugacities are taken from Prausnitz et al.22
A computer program that was written in the C++ language
was used to apply the developed algorithm in this work. The
program is compiled using Visual C++ Version 5.0. The
required numbers of outer-loop iterations and CPU times to
converge the algorithm for the examples are given in Table 2.
These examples are executed on a personal computer that is
equipped with a 2.4 GHz Pentium 4 processor.
Two-Phase Reactive Distillation Columns. (A) Example
1: Methyl Acetate Production. Methyl acetate is used as a
solvent for lacquers, celluloses, paints, resins, coatings, and
perfumes. It is also used as an intermediate chemical in synthetic
flavors and artificial leather manufacturing. Methyl acetate can
be produced via the esterification of methanol with acetic acid
in the liquid phase in the presence of an acidic catalyst. The
equilibrium reaction may be written as follows:
H+

CH3COOH + CH3OH 798 CH3COOCH3 + H2O (36)


The catalyst can be either a homogeneous catalyst, such as
sulfuric acid,23 or a heterogeneous catalyst, such as an acidic
ion-exchange resin.24,25 Conventional production of high-purity
methyl acetate is very difficult, because of the chemical
equilibrium limitations in the reactor and also the presence of
methanol-methyl acetate and water-methyl acetate minimumboiling azeotropes in the separation section of the process. The
Eastman Kodak Companys conventional process for the
production of methyl acetate is a complex process that consists
of one reactor followed by nine separation units.2,3 The reactive
distillation process first developed and commercialized by the

Ind. Eng. Chem. Res., Vol. 45, No. 17, 2006 6013
Table 1. Column Specifications for Two- and Three-Phase Reactive Distillation Examples
Feed Conditions
Compositiona
component

Column Specifications

FI

FII

acetic acid
methanol
methyl acetate
water

1.0
0.0
0.0
0.0

0.0
1.0
0.0
0.0

iso-butylene
1-butene
ethanol
ETBE

0.073
0.545
0.091
0.291

F6: 100

butanol
acetic acid
butyl acetate
water

0.18
0.17
0.32
0.33

F23: 18.8

butanol
acetic acid
butyl acetate
water

0.18
0.17
0.32
0.33

F12: 18.8

hexanol
acetic acid
hexyl acetate
water

1.0
0.0
0.0
0.0

cyclohexane
cyclohexene
cyclohexanol
water

0.1
0.4
0.0
0.5

0.0
1.0
0.0
0.0

rate (kmol/h)
F8: 100
F21: 100

F8: 50
F14: 50

F6: 100

conditions

number of stages, Ns

Example 1 (see refs 23-27 and 29)


saturated liquid
27
P: 0.1013 (MPa)
saturated liquid
P: 0.1013 (MPa)
Example 2 (see refs 30 and 31)
liquid
10
T: 303 K
P: 0.95 (MPa)
Example 3a (see refs 5-8, 32, 33)
liquid
47
T: 360 K
P: 0.1013 (MPa)
Example 3b (see refs 5-8,32,33)
liquid
25
T: 360 K
P: 0.1013 (MPa)
Example 4 (see ref 4)
saturated liquid
23
P: 0.04 (MPa)
saturated liquid
P: 0.04 (MPa)
Example 5 (see refs 9, 10, 34)
saturated liquid
20
P: 0.1013 (MPa)

reactive stages

reflux ratio

P (MPa)

B (kmol/h)

8-20

R: 2.2

0.1013

100

3-5

R: 4

0.95

21-27

RI: 99
RII: 0 (aq.)

0.1013

8.74

11-15

RI: 99
RII: 0 (aq.)

0.1013

8.74

8-13

RI: total
RII: 0 (aq.)

0.04

50

2-11

RI: 9 (org.)
RII: 3.6

0.1013

40

38

FI, first feed from top; FII, second feed from top.

same company can produce high-purity methyl acetate with only


one reactive distillation column.23 An Eastman Kodak plant in
Tennessee produces 180 000 tons/yr of high-purity methyl
acetate with only a single reactive distillation column.3 This
process is an example of the success of the reactive distillation
technology, and it has been studied by many authors.23-27
The column specification for Example 1 is given in Table 1,
and this information is similar to the specification of the pilotplant scale column studied by Bessling et al.24 For this example,
the NRTL activity model with parameters taken from Gmehling
et al.28 is used to calculate the liquid phase thermodynamic
properties (see Table A1 in the Appendix). The esterification
reaction for producing methyl acetate over a heterogeneous
catalyst (using Amberlyst 15W, which is an ion-exchange resin)
was studied by Song et al.29 The activity-based reaction
equilibrium constant for the esterification reaction is taken from
Song et al.29 and is given by

Keq ) 2.32 exp

(782.98
T )

(37)

where the temperature T is given in Kelvin.


For this example, the algorithm converged within 22 iterations
and a CPU time of 16.7 s, as given in Table 2. The simulation
results are presented in Figures 4-6. Figures 4 and 5 show the
column temperature and liquid-phase composition profiles,
respectively, for a reflux ratio of 2.2. Figure 5 shows that a
high-purity distillate product (97.12 mol % methyl acetate) can
be obtained at this reflux ratio, whereas virtually no methyl
acetate is lost through the bottom product. Figure 6 compares

Table 2. Number of Iterations and Computational Time for


Examples of Table 1

example

outer-loop iteration

CPU time (s)a

1
2
3a
3b
4
5

22
5
13
6
6
21

16.7
18.2
57.6
68.4
17.8
88.6

Using a 2.4 GHz Pentium 4 processor.

the simulated acetic acid conversion in this work with the


experimental data obtained by Bessling et al.24 It can be
observed that the simulated acetic acid conversion shows good
agreement with the experimental data obtained by Bessling et
al.24 The results show that the conversion increases rapidly as
the reflux ratio increases until it reaches a maximum value of
97.12% at the optimum reflux ratio of 2.2. A similar experimental optimum reflux ratio was observed by Bessling et al.;
this value corresponded to the maximum experimental conversion value of 96.95%. In Figure 6, the deviation between
simulated acetic acid conversion and experimental data at higher
reflux ratios is due to a reduction in the liquid residence time
in the experimental column, where the assumption of chemical
equilibrium reaction is no longer a very good assumption.24,25
Nevertheless, the chemical reaction equilibrium assumption
provides an excellent prediction in the range of optimal reflux
ratio.
(B) Example 2: Ethyl tert-Butyl Ether (ETBE) Production. ETBE is an ether that is used as an oxygenate additive
for enhancing the octane number of gasoline fuel. Recently,

6014

Ind. Eng. Chem. Res., Vol. 45, No. 17, 2006

Figure 4. Stage temperature profile in Example 1 for R ) 2.2; feeds are


introduced to stages 8 and 21; reactive stages include stages 8-20.

Figure 7. Stage temperature profile in Example 2; feed is introduced to


stage 6; reactive stages include stages 3-5.

Figure 8. Liquid-phase composition profile in Example 2.


Figure 5. Liquid-phase composition profile in Example 1 (R ) 2.2).

occurs in the liquid phase. The reaction extent is only 84.7%


from a stoichiometric mixture of reactants at 70 C. The
equilibrium conditions of this reaction were studied by Jensen
and Datta.30 The reaction equilibrium constant for ETBE
reaction (eq 39) resulting from this study was estimated as
follows:

4060.59
- 2.89055 ln T T
0.0191544T + 5.28586 10-5T2 - 5.32977 10-8T3
(39)

ln Keq ) 10.387 +

Figure 6. Comparison of the simulated acetic acid conversion in this work


with the experimental data obtained by Bessling et al.24 in Example 1.

much attention has been given to ETBE production, because


of its higher-octane-enhancing properties and lesser fuel vaporization loss, which is due to its lower volatility. Also, this
low vapor pressure of ETBE reduces the emission of volatile
organic compounds of fuel. Therefore, ETBE is considered an
environmentally friendly fuel additive. ETBE is the product of
the etherification reaction of isobutylene with ethanol over an
acidic catalyst, such as an acidic ion-exchange resin (e.g.,
Amberlyst 15).

(CH3)2CdCH2 + C2H5OH T (CH3)3COC2H5

(38)

The reaction is an exothermic equilibrium limited reaction that

where T is given in Kelvin.


ETBE can be produced in a reactive distillation column to
eliminate the equilibrium limitation and attain a high conversion
of olefin feed and produce high-purity product. Example 2 is
taken from Sneesby et al.31 The thermodynamic properties for
the liquid phase are calculated using the UNIQUAC activity
model (see Table A2 in the Appendix). For this example, the
algorithm is converged within five iterations and a CPU time
of 18.2 s. The simulation results for this example are presented
in Figures 7 and 8, for temperature and liquid composition
profile, respectively. The composition profile shows that the
bottom stream is a high-purity (95.25 mol % ETBE) product.
A high isobutylene conversion of 99.96 mol % is obtained by
this reactive distillation column.
Three-Phase Distillation Columns. (A) Example 3: Heterogeneous Azeotropic Reactive Distillation Column, Butyl
Acetate Production. Butyl acetate is a widely used solvent in
coatings and paints industries. It is also used in leather treatment

Ind. Eng. Chem. Res., Vol. 45, No. 17, 2006 6015

Figure 9. Comparison of simulated temperature profile in this work with


the experimental data obtained by Hanika et al.6 in Example 3a; feed is
introduced to stage 23; reactive stages include stages 21-27.

Figure 10. Organic liquid-phase composition profile in Example 3a.

and as an intermediate solvent in the manufacture of adhesives.


Butyl acetate can be produced via the liquid-phase reaction of
butanol and acetic acid in the presence of a suitable acidic
catalyst. The esterification reaction is given by
H+

CH3COOH + C4H9OH 798CH3COOC4H9 + H2O (40)


This reaction, in the presence of an ion-exchange resin as a
catalyst, has been studied by Gangadwala et al.,32 and the
activity-based reaction equilibrium constant is given by

Keq ) 3.8207 exp

(430.803
T )

(41)

where T is given in Kelvin.


The production of butyl acetate via reactive distillation is an
example of a heterogeneous azeotropic reactive distillation
process. In this process, a minimum-boiling ternary azeotrop
mixture (bp 90 C) between butyl acetate, butanol, and water
is formed on the top stage of the column.7,33 The azeotropic
vapor mixture splits in two organic and aqueous liquid phases
in the decanter when it is condensed in the condenser. The
aqueous liquid phase that contains mainly water is totally
withdrawn from the column and the major portion of the organic
liquid phase is refluxed to the column. The bottom stream is a
high-purity butyl acetate product. Hanika et al.6 studied the
production of butyl acetate experimentally, using a packed
reactive distillation column. Example 3a in Table 1 gives
specifications for this column. The number of theoretical stages
for the column reported in Table 1 was obtained by the authors,6
using the height and the number of theoretical stages per meter
(NTSM) of packing. The reactive section is located in the middle
of the column, and the feed is introduced into the reactive
section. The UNIQUAC model, with parameters taken from
Venimadhavan et al.,7 is used to model the liquid-phase
thermodynamic properties (see Table A3 in the Appendix). The
simulation results show the formation of two liquid phases on
the first stage (condenser-decanter). Figure 9 shows the
temperature profile in the column. The figure shows that the
predicted temperature profile is in very good agreement with
the experimental temperature profile obtained by Hanika et al.6
Figure 10 presents the liquid-phase mole fraction profile in the
column (organic liquid for the first stage). Hanika et al.6
suggested that at least 20 theoretical stages in the stripping
section of the column are necessary to accomplish a good
separation efficiency for separating acetic acid from butyl
acetate. However, as can be seen from the simulation results in
Figure 10, the liquid phase becomes almost-pure butyl acetate

Figure 11. Organic liquid-phase composition profile in Example 3b.


Table 3. Distillate and Bottom Product Liquid Compositions for
Example 3a (47-Stage Column) and Example 3b (25-Stage Column)
Liquid Mole Fraction
Example 3a
component

Example 3b

aqueous organic bottom aqueous organic bottom

butanol
acetic acid
butyl acetate
water

0.0134
0.0223
0.0012
0.9631

0.3539
0.0686
0.2037
0.3738

0.0000
0.0000
1.0000
0.0000

0.0123
0.0203
0.0012
0.9662

0.3449
0.0626
0.2359
0.3566

0.0001
0.0000
0.9999
0.0000

flow rate (mol/h)

9.1138

0.9462 8.7400

9.1355

0.9245 8.7400

in the stripping section of the column immediately after the


reactive zone. This observation suggests the idea that a large
number of stages in the stripping section of the column for
achieving a high-purity product is not actually necessary. To
examine this idea, a smaller column with total of 25 theoretical
stages (Example 3b in Table 1) is simulated. Under the same
operating conditions as the previous 47-stage column, the
composition profile for Example 3b is shown in Figure 11 and
Table 3. It is noted, from the results, that the same high-purity
product is achieved for the column with 25 stages as was
achieved with the 47-stage column. The computation time and
number of iterations for Examples 3a and 3b are given in Table
2.
(B) Example 4: Heterogeneous Azeotropic Reactive
Distillation Column, Hexyl Acetate Production. Hexyl acetate
is a fruity-smelling substance that is used as a flavoring agent
in food industries and as aroma in perfume manufacturing.
Hexyl acetate is produced via the liquid-phase esterification
reaction of hexanol with acetic acid over a strong acid catalyst.
The liquid-phase reversible reaction is given by
H+

CH3COOH + C6H13OH 798 CH3COOC6H13 + H2O (42)

6016

Ind. Eng. Chem. Res., Vol. 45, No. 17, 2006

12, because of the variation of the concentrations of hexyl


acetate and acetic acid (nbp 118.1 C), which is introduced as
feed on stage 14. Figure 13 shows the organic liquid-phase
composition profile. This figure shows that a high-purity bottom
product (99.91 mol % hexyl acetate) can be produced by this
process. This example converged within six iterations and a CPU
time of 17.8 s.
(C) Example 5: Three-Phase Reactive Distillation Column, Cyclohexanol Production. Cyclohexanol is a saturated
alicyclic alcohol that is used in the production of intermediates
for nylons, plasticizers, pesticides, and detergents. Cyclohexanol
can be produced via the hydration of cyclohexene with water
over a suitable catalyst. The hydration reaction is given by
cat

cyclohexene + water 798 cyclohexanol


Figure 12. Stage temperature profile in Example 4. Feeds are introduced
to stages 7 and 14; reactive stages include stages 8-13.

Figure 13. Organic liquid-phase composition profile in Example 4.

This esterification reaction in the presence of an ion-exchange


resin (Amberlyst CSP2) as a catalyst has been studied by
Schmitt et al.4 They estimated an activity-based reaction
equilibrium constant that is independent of temperature, given
by

Keq ) 31.9281

(43)

This liquid-phase equilibrium reaction can be performed in a


reactive distillation column. The column specifications for the
production of hexyl acetate by the reactive distillation process
are given in Example 4 of Table 1. The NRTL model with
parameters taken from Schmitt et al.4 is used to calculate the
liquid phase thermodynamic properties (see Table A4 in the
Appendix).
The simulation results of this example show the formation
of two liquid phases on the first stage (condenser-decanter).
The aqueous phase contains mainly water (99.92 mol % H2O)
and is withdrawn completely from the decanter. Figure 12 shows
the stage temperature profile in the column. This figure shows
that the stage temperature decreases after stage 8, where the
hexanol (nbp 157.2 C) feed stream is introduced to the column.
After stage 10, the stage temperature increases as the concentration of hexanol decreases (by consumption in the chemical
reaction) and the concentration of hexyl acetate, which is a
compound with a higher boiling point (nbp 171.3 C), increases
via production in the chemical reaction (Figure 13). Another
decline and climb in stage temperature is observed after stage

(44)

The industrial process for the production of cyclohexanol, which


was developed by the Asahi Chemical Co., consists of a slurry
reactor, followed by a decanter vessel and a distillation column.10
The hydration reaction occurs in the slurry reactor, where
cyclohexene and water come into contact with a very fine
catalyst (zeolite catalyst, H-ZSM 5). Subsequently, the reaction
product is separated into organic and aqueous phases in the
decanter. The catalyst is recovered in the liquid phase and
recycled to the reactor, and the organic phase that contains the
cyclohexanol product is sent to the distillation section for
purification.
The feasibility of cyclohexanol production via a reactive
distillation process was examined by Steyer et al.10 and Qi et
al.,9 using batch reactive distillation experiments and residue
curve map studies. Their residue curve maps study shows that
the production of high-purity cyclohexanol via a reactive
distillation process is conceptually possible. However, as they
indicated, the simulation of a multistage reactive column for
this process involves many complexities, because of the
simultaneous presence of reaction, distillation, and liquid-phase
splitting in the system.
Using the algorithm developed in the present paper, a threephase reactive distillation column for the production of cyclohexanol is simulated. The column specifications are given in
Table 1. The feed is a mixture of water and hydrocarbon
compounds. The hydrocarbon mixture is considered to be a
mixture of cyclohexene as the reactive component and cyclohexane as an inert impurity that comes from an upstream unit.
The UNIQUAC model with parameters taken from Gmehling
et al.28 is used to model the liquid-phase thermodynamic
properties (see Table A5 in the Appendix). The liquid-phase
hydration of cyclohexene catalyzed by a strong acid ionexchange resin was studied by Panneman and Beenackers.34 The
activity-based reaction equilibrium constant for this reaction is
given by

Keq ) 2.37 10-5 exp

(3636.757
)
T

(45)

where T is given in Kelvin.


Simulation results for this example are presented in Figures
14-18. The computation for this example converged within 21
iterations and a CPU time of 88.6 s. Figure 14 shows the vapor
and liquid profiles in the column. As it can be seen, two liquid
(organic and aqueous) phases are present from stage 1 through
stage 18. The second liquid (aqueous) phase does not exist on
stages 19 and 20. The organic liquid phase leaves the bottom
of the column as a high-purity cyclohexanol product. This phase
is partially (10% of organic liquid) withdrawn from the top of

Ind. Eng. Chem. Res., Vol. 45, No. 17, 2006 6017

Figure 14. Vapor and liquid flow rates in Example 5.

Figure 17. Effect of the organic liquid reflux ratio (RI) on the cyclohexene
conversion and bottom product organic-phase cyclohexanol purity in
Example 5.

Figure 15. Organic-phase (liquid I, LI) composition profile in Example 5.


Figure 18. Effect of the organic liquid reflux ratio (RI) on aqueous-phase
(liquid II, LII) fraction profile in Example 5.

Figure 16. Stage temperature profile in Example 5.

the column as a distillate product that contains a high concentration of cyclohexane. The organic liquid phase composition
profile is shown in Figure 15. The aqueous phase contains
mainly water. The temperature profile is shown in Figure 16.
The cyclohexene conversion, cyclohexanol purity in the
bottom product, and the phase pattern in the column are all
dependent on the operating conditions of reactive distillation
column. The effect of the organic phase reflux ratio on the
cyclohexene conversion and the cyclohexanol purity of organic
phase in the bottom product is examined and shown in Figure
17. The cyclohexene conversion increases as the organic phase
reflux ratio increases until it reaches a maximum (98%
conversion) near an organic reflux ratio of RI ) 9. It then begins
to decrease with further increases in the reflux ratio. The
cyclohexanol purity of the bottom product does not show a
major change with increases in the organic reflux ratio until it
reaches an organic reflux ratio of RI 0.5, where the phase
fraction of aqueous liquid approaches to zero on the last stage.

It then increases with the disappearance of the aqueous liquid


phase on the last stage with further increases in reflux ratio until
it reaches a maximum at RI 9, which is the same reflux ratio
for maximum cyclohexene conversion. Therefore, there is an
optimum reflux ratio in which both conversion and purity are
at their maximum. It is important, from an operational point of
view, to keep the reflux ratio near its optimum value for high
efficiency of the reactive distillation column. It is also important
to know that an increase in reflux ratio does not always increase
the purity and conversion. Hence, a specific consideration should
be given in choosing a control strategy for this type of column.
The phase pattern within the column also is dependent on
the organic reflux ratio. Figure 18 shows the aqueous liquid
(LII) phase fraction profile for different organic liquid reflux
ratios. As it can be seen, two liquid phases are present on all
stages for an organic reflux ratio of RI ) 0.5. The aqueous liquid
phase begins to disappear gradually from the bottom stages of
the column as the organic phase reflux ratio increases.
Conclusions
An algorithm for the simulation of two- and three-phase
multistage reactive distillation columns was developed. In this
algorithm, the phase stability, phase equilibrium, and chemical
reaction equilibrium calculations in three-phase reactive distillation columns are performed simultaneously. A prior knowledge
of the phase pattern in the column is not required. The resulting
nonlinear model equations in the reactive distillation system are
solved using an inside-out method.
The algorithm was used to simulate a variety of two- and
three-phase reactive distillation columns. The results of five

6018

Ind. Eng. Chem. Res., Vol. 45, No. 17, 2006

different reactive distillation examples are presented in this work.


The simulation results of two examples were compared with
the available experimental data in the literature. Good agreement
was observed between the simulated results and the experimental
data. The algorithm was efficient and robust and converged
within a reasonable number of iterations and computational time
for all the examples.
The algorithm was used to simulate a multistage three-phase
reactive distillation for the production of cyclohexanol. It was
observed that a high-purity cyclohexanol product and a high
reactant (cyclohexene) conversion can be obtained using the
reactive distillation process. The high-purity product was
obtained when a single liquid phase existed on the last stage.
The cyclohexanol reactive column performance was studied
under different operating conditions. The process exhibited a
maximum conversion and product purity at a specific optimum
reflux ratio. The ability of the algorithm in determining the phase
pattern within the reactive distillation column was shown in
the cyclohexanol reactive column by changing the column
operating conditions.
Acknowledgment
The financial support for this work, provided by the Natural
Science and Engineering Research Council (NSERC) of Canada,
is greatly appreciated.

Table A3. UNIQUAC Binary Parameters (aij(K)) and UNIQUAC


Structural Parameters for Example 3
aij(K)

i ) butanol
i ) acetic acid
i ) butyl acetate
i ) water

j)
butanol

j)
acetic acid

j)
butyl acetate

j)
water

0
-66.308
12.398
292.444

74.619
0
358.409
265.662

41.532
-150.178
0
232.222

34.223
-172.902
345.062
0

component

butanol
acetic acid
butyl acetate
water

3.454
2.202
4.827
0.920

3.052
2.072
4.196
1.400

3.052
2.072
4.196
1.400

Table A4. NRTL Binary Parameters (aij(K) and ij)a and NRTL rij
Parameters for Example 4
j)
hexanol
i ) hexanol
i ) acetic acid
i ) hexyl acetate
i ) water

j)
water

-1049.70
-1489.92
0
3545.58

-690.50
119.03
998.70
0

0
0
-4.065
18.510

0
0
-5.456
1.805

3.241
4.290
0
-1.748

3.466
-0.746
-1.315
0

Combination i-j

Table A1. NRTL Binary Parameters (aij (K))a and NRTL rij
Parameters for Example 1
aij(K)

j)
acetic acid

j)
methanol

j)
methyl acetate

j)
water

0
-78.839
733.371
1360.576

54.402
0
309.351
560.141

-398.460
76.818
0
762.181

-693.496
-203.072
240.178
0

Combination i-j
component i

component j

Rij ) Rji

acetic acid
acetic acid
acetic acid
methanol
methanol
methyl acetate

methanol
methyl acetate
water
methyl acetate
water
water

0.3067
0.3026
0.1463
0.2968
0.3004
0.2152

Used in eq A.2 in the Appendix.

Table A2. UNIQUAC Binary Parameters (aij(K)) and UNIQUAC


Structural Parameters for Example 2

component i

component j

Rij ) Rji

hexanol
hexanol
hexanol
acetic acid
acetic acid
hexyl acetate

acetic acid
hexyl acetate
water
hexyl acetate
water
water

0.3
0.3
0.3
0.3
0.3
0.2

Used in eq A.3 in the Appendix.

Table A5. UNIQUAC Binary Parameters (aij(K)) and UNIQUAC


Structural Parameters for Example 5
aij(K)
j)
j)
j)
cyclohexane cyclohexene cyclohexanol

j)
water

-49.121
0
76.514
466.350

1247.300
1024.100
159.289
0

i ) cyclohexane
i ) cyclohexene
i ) cyclohexanol
i ) water

0
60.813
-150.907
540.360

component

4.046
3.814
4.349
0.920

3.240
3.027
3.512
1.400

3.240
3.027
1.780
1.000

The NRTL equation is given as


Nc

j ) iso-butene

j ) 1-butene

j ) ethanol

j ) ETBE

0
-23.894
436.034
39.215

24.245
0
404.721
42.130

-46.937
-26.930
0
-102.322

-21.484
-20.041
424.521
0

component

iso-butene
1-butene
ethanol
ETBE

2.92
2.92
2.11
5.86

2.68
2.56
1.97
4.94

2.68
2.56
0.92
4.94

448.368
-53.743
0
128.476

cyclohexane
cyclohexene
cyclohexanol
water

aij(K)
i ) iso-butene
i ) 1-butene
i ) ethanol
i ) ETBE

j)
hexyl acetate

ij

i ) hexanol
i ) acetic acid
i ) hexyl acetate
i ) water

The NRTL and UNIQUAC parameters are given in Tables


A1, A2, A3, A4, and A5 for Examples 1, 2, 3, 4, and 5,
respectively.

aij(K)
579.91
0
2184.61
-86.26

0
-269.12
1522.47
-3501.50

Appendix: NRTL and UNIQUAC Parameters

i ) acetic acid
i ) methanol
i ) methyl acetate
i ) water

j)
acetic acid

ln i )

j)1

( )
Nc

jiGjixj

Nc

Glixl

l)1

Nc

xjGij

xrrjGrj

r)1

ij
N
N
j)1
Gljxl
Gljxl

l)1
l)1
c

(A.1)

where

ij )

aij
T

(A.2)

Ind. Eng. Chem. Res., Vol. 45, No. 17, 2006 6019

or

ij )

aij
+ ij
T

(A.3)

and

Gji ) exp(-Rjiji)

(Rij ) Rji)

(A.4)

The UNIQUAC equation is given as

ln i ) ln

( ) ( ) ( ) ( )
i
xi

qi ln

+ li -

Nc

qi ln(

j)1

i
xi

Nc

xjlj -

j)1
Nc

jij

N
j)1
kkj
k)1

jji) + qi - qi

(A.5)

where

z
li ) (ri - qi) - (ri - 1)
2
i )

(for z ) 10)

(A.6)

rixi
(A.7)

Nc

rjxj

j)1
i )

Greek Symbols

qixi
(A.8)

Nc

qjxj

j)1
i )

qixi
(A.9)

Nc

qjxj

j)1
ij ) exp

( )
-aij
T

Hf ) feed enthalpy (kJ)


K ) equilibrium ratio
Keq ) chemical reaction equilibrium constant
l ) parameter given by eq A.6
L ) liquid flow rate (kmol/h)
m ) total molar flow rate (kmol/h)
n ) molar flow rate (kmol/h)
Nc ) number of components
Ns ) number of stages
Nr ) number of reactions
Nrs ) number of reactive stages
q ) UNIQUAC parameter in eqs A.5, A.6, A.8, and A.9
Q ) stage energy input (kJ)
r ) UNIQUAC parameter in eqs A.6 and A.7
R ) reflux ratio
R ) universal gas constant (kJ kmol-1 K-1)
Ri ) generation of component i due to reaction(s) (kmol)
RL ) liquid reflux rate (kmol/h)
RVL ) vapor liquid ratio
T ) temperature (K)
V ) vapor flow rate (kmol/h)
WL ) liquid side withdrawal (kmol/h)
WV ) vapor side withdrawal (kmol/h)
wr ) side-withdrawal ratio
x ) mole fraction or stage independent variable vector
X ) independent variables vector
z ) feed mole fraction

(A.10)

Nomenclature
a ) coefficient in eq 32 or in eqs A.2, A.3, and A.10
a i ) activity of component i in the mixture
b ) coefficient in eq 32
B ) bottom product molar rate (kmol/h)
BL ) bottom liquid product (kmol/h)
c ) coefficient in eq 32
-1 kmol-1)
Cig
p ) ideal-gas heat capacity (kJ K
d ) coefficient in eq 33
DL ) distillate liquid product (kmol/h)
DV ) distillate vapor product (kmol/h)
e ) coefficient in eq 33
f ) model equation functions or coefficient in eq 33
fi ) fugacity of component i in the mixture
F ) model equations vector
F ) feed molar rate (kmol/h)
G ) NRTL parameter given by eq A.4
hhi ) partial molar enthalpy of component i in the mixture (kJ/
kmol)
hig
i ) ideal gas molar enthalpy of component i (kJ/kmol)
hhRi ) partial molar residual enthalpy of component i in the
mixture (kJ/kmol)

R ) phase fraction or parameter in eq A.4


) parameter in eq A.3
hi,f298 ) standard heat of formation of component i (1 atm,
25 C)
) discrepancy or function residual
) UNIQUAC segment fraction given by eq A.7
i ) fugacity coefficient of component i in the mixture
i ) activity coefficient of component i in the mixture
) stoichiometric coefficient
) number of phases
) phase stability variable
) UNIQUAC area fraction given by eqs A.8 and A.9
) component phase fraction
) stage parameter or interaction parameter given by eqs A.2,
A.3, and A.10
) extent of reaction
Literature Cited
(1) Backhaus, A. A. Continuous Processes for the Manufacture of Esters.
U.S. Patent No. 1,400,849, 1921.
(2) Taylor, R.; Krishna, R. Modeling Reactive Distillation. Chem. Eng.
Sci. 2000, 55, 5183.
(3) Sundmacher, K., Kienle, A., Eds. ReactiVe Distillation: Status and
Future Directions; Wiley-VCH Verlag GmbH & Co.: Weinheim, Germany, 2001.
(4) Schmitt, M.; Hasse, H.; Althaus, K.; Schoenmakers, H.; Gotze, L.;
Moritz, P. Synthesis of n-Hexyl Acetate by Reactive Distillation. Chem.
Eng. Process. 2004, 43, 397.
(5) Zhicai, Y.; Xianbao, C.; Jing, G. Esterification-Distillation of Butanol
and Acetic Acid. Chem. Eng. Sci. 1998, 53, 2081.
(6) Hanika, J.; Kolena, J.; Smejkal, Q. Butyl Acetate via Reactive
DistillationsModeling and Experiment. Chem. Eng. Sci. 1999, 54, 5205.
(7) Venimadhavan, G.; Malone, M. F.; Doherty, M. F. A Novel Distillate
Policy for Batch Reactive Distillation with Application to the Production
of Butyl Acetate. Ind. Eng. Chem. Res. 1999, 38, 714.
(8) Gangadwala, J.; Kienle, A.; Stein, E.; Mahajani, S. Production of
Butyl Acetate by Catalytic Distillation: Process Design Studies. Ind. Eng.
Chem. Res. 2004, 43, 136.

6020

Ind. Eng. Chem. Res., Vol. 45, No. 17, 2006

(9) Qi, Z.; Kolah, A.; Sundmacher, K. Residue Carve Maps for Reactive
Distillation Systems with Liquid-Phase Splitting. Chem. Eng. Sci. 2002,
57, 163.
(10) Steyer, F.; Qi, Z.; Sundmacher, K. Synthesis of Cyclohexanol by
Three-Phase Reactive Distillation: Influence of Kinetics on Phase Equilibria.
Chem. Eng. Sci. 2002, 57, 1511.
(11) Qi, Z.; Sundmacher, K. Bifurcation Analysis of Reactive Distillation
Systems with Liquid-Phase Splitting. Comput. Chem. Eng. 2002, 26, 1459.
(12) Gumus, Z. H.; Ciric, A. R. Reactive Distillation Column Design
with Vapor/Liquid/Liquid Equilibria. Comput. Chem. Eng. 1997, 21, S983.
(13) Khaledi, R.; Bishnoi, P. R. A Method for Steady-State Simulation
of Multistage Three-Phase Separation Columns. Ind. Eng. Chem. Res. 2005,
44, 6845.
(14) Gupta, A. K.; Bishnoi, P. R.; Kalogerakis, N. A Method for the
Simultaneous Phase Equilibria and Stability Calculations for Multiphase
Reacting and Non-Reacting Systems. Fluid Phase Equilib. 1991, 63, 65.
(15) Khaledi, R. Simulation of Multiphase Reactive and Non-Reactive
Separation Processes. Ph.D. Thesis, University of Calgary, Calgary, Alberta,
Canada, 2005.
(16) Boston, J. F.; Sullivan, S. L. A New Class of Solution Methods
for Multicomponent, Multistage Separation Processes. Can. J. Chem. Eng.
1974, 52, 52.
(17) Naphthali, L. M.; Sandholm, D. P. Multicomponent Separation
Calculations by Linearization. AICHE J. 1971, 17, 148.
(18) Rheinboldt, W.; Burkardt, J. A Program for a Locally Parametrized
Continuation Process. ACM Trans. Math. Soft. 1983, 9, 236.
(19) Powell, M. J. D. A Hybrid Method for Non-Linear Equations. In
Numerical Methods for Nonlinear Algebraic Equations; Rabinowitz, P., Ed.;
Gordon and Breach: New York, 1970.
(20) Trebble, M. A. A Preliminary Evaluation of Two and Three Phase
Flash Initiation Procedures. Fluid Phase Equilib. 1989, 53, 113.
(21) Peng, D. Y.; Robinson, D. B. A New Two-Constant Equation of
State. Ind. Eng. Chem. Fundam. 1976, 15, 59.
(22) Prausnitz, J. M.; Anderson, T. F.; Grens, E. A.; Eckerts, C. A.;
Hsieh, R.; OConnell, J. P. Computer Calculations for Multicomponent
Vapor-Liquid and Liquid-Liquid Equilibria; Prentice Hall: Englewood
Cliffs, NJ, 1980.
(23) Agreda, V. H.; Partin, L. R.; Heise, W. H. High-Purity Methyl
Acetate via Reactive Distillation. Chem. Eng. Prog. 1990, 86, 40.

(24) Bessling, B.; Loning, J. M.; Ohligschlager, A.; Schembecker, G.;


Sundmacher, K. Investigations on the Synthesis of Methyl Acetate in a
Heterogeneous Reactive Distillation Process. Chem. Eng. Technol. 1998,
21, 393.
(25) Popken, T.; Steinigeweg, S.; Gmehling, J. Synthesis and Hydrolysis
of Methyl Acetate by Reactive Distillation Using Structured Catalytic
Packings: Experiments and Simulation. Ind. Eng. Chem. Res. 2001, 40,
1566.
(26) Barbosa, D.; Doherty, M. F. Design and Minimum-Reflux Calculations for Double-Feed Multicomponent Reactive Distillation Columns.
Chem. Eng. Sci. 1988, 43, 2377.
(27) Huss, R. S.; Chen, F.; Malone, M. F.; Doherty, M. F. Reactive
Distillation for Methyl Acetate Production. Comput. Chem. Eng. 2003, 27,
1855.
(28) Gmehling, J.; Onken, U. Vapor-Liquid Equilibrium Data Collection; DECHEMA Chemistry Data Series, Vol. 1, Part 1; 1977.
(29) Song, W.; Venimadhavan, G.; Manning, J. M.; Malone, M. F.;
Doherty, M. F. Measurement of Residue Carve Maps and Heterogeneous
Kinetics in Methyl Acetate Synthesis. Ind. Eng. Chem. Res. 1998, 37, 1917.
(30) 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.
(31) Sneesby, M. G.; Tade, M. O.; Datta, R.; Smith, T. N. ETBE
Synthesis via Reactive Distillation. 1. Steady-State Simulation and Design
Aspects. Ind. Eng. Chem. Res. 1997, 36, 1855.
(32) Gangadwala, J.; Mankar, S.; Mahajani, S. Esterification of Acetic
Acid with Butanol in the Presence of Ion-Exchange Resins as Catalysts.
Ind. Eng. Chem. Res. 2003, 42, 2146.
(33) Loning, S.; Horst, C.; Hoffmann, U. Theoretical Investigations on
the Quaternary System n-Butanol, Butyl Acetate, Acetic Acid and Water.
Chem. Eng. Technol. 2000, 23, 789.
(34) Panneman, H. J.; Beenackers, A. A. C. M. Solvent Effects in the
Liquid-Phase Hydration of Cyclohexene Catalyzed by a Macroporous Strong
Acid Ion Exchange Resin. Chem. Eng. Sci. 1992, 47, 2635.

ReceiVed for reView December 12, 2005


ReVised manuscript receiVed May 10, 2006
Accepted June 19, 2006
IE051384A

Anda mungkin juga menyukai