You are on page 1of 15

Ind. Eng. Chem. Res.

2004, 43, 2493-2507 2493

PROCESS DESIGN AND CONTROL

Quantitative Comparison of Reactive Distillation with Conventional


Multiunit Reactor/Column/Recycle Systems for Different Chemical
Equilibrium Constants
Devrim B. Kaymak and William L. Luyben*
Process Modeling and Control Center, Department of Chemical Engineering, Lehigh University,
Bethlehem, Pennsylvania 18015

Recent advances in technologies based on process intensification offer different design alternatives
for chemical processes including reaction and separation units. An excellent example of process
intensification is reactive distillation, which combines reaction and separation units in one piece
of equipment. This paper studies the steady-state economic optimum design of a chemical process
with a generic exothermic reversible reaction A + B T C + D for two different process
flowsheets: a conventional multiunit reactor/separator/recycle structure and a reactive distillation
column. Each system is optimized in terms of the total annual cost for a wide range of chemical
equilibrium constants, KEQ. In the conventional system, the design optimization variables include
the reactor temperature, reactor size, and recycle flow rate. In the reactive distillation system,
the design optimization variables include the pressure, number of reactive trays, and number
of total trays. The two systems are designed for identical feeds and identical products. Results
show that reactive distillation is significantly less expensive (by a factor of up to 3) than the
conventional process for all values of the chemical equilibrium constant.

1. Introduction tants remain in the column. This means that the


products should be lighter or heavier than the reactants.
If one of the most important operations in chemical The temperatures of reaction and separation should be
engineering is separation, with distillation columns as similar because both of these operations occur in the
its most popular process unit, reactors are the other same unit simultaneously. The reaction should be fairly
vital parts of the chemical industry. Since reactor
fast, so that the liquid holdups or amounts of catalyst
effluents fail to meet specification criteria because of
required on each reactive tray are feasible in light of
the presence of unconverted materials, reactors followed
hydraulic limitations (reasonable tray liquid heights).
by several distillation columns with recycle streams
Heats of reaction cannot be too large because of their
back to the reaction section are common in industry.
impact on the vapor and liquid flow rates throughout
Economic and environmental considerations have
the reactive zone. For example, if the reaction is
encouraged industry to focus on technologies based on
exothermic, the liquid flow rates decrease from tray to
process intensification. This is an area of growing
tray down through the reactive section of the column,
interest that is defined as any chemical engineering
and the vapor flow rates increase from tray to tray up
development that leads to a substantially smaller and
through the reactive section. If the heat of reaction is
more energy-efficient technology.1 An excellent example
of process intensification is reactive distillation, which large, hydraulic problems such as weeping or flooding
combines reaction and separation units in a single can occur.
vessel. Reactive distillation can, in some systems, If reactive distillation is suitable for a chemical
provide an alternative to conventional multiunit flow- system, then this process provides several advantages
sheets, which typically include a reactor followed by a over the conventional flowsheet. Because products are
separation section with recycles back to the reaction continuously removed from the reactive zone, the con-
section. version can increase significantly. The selectivity can
Although reactive distillation might be an attractive also be increased because of the continuous separation
alternative to the conventional multiunit processes, it between reactants and products. A reduction of total
can be effective for only a fairly small class of chemical investment and operating costs is usually achieved
systems because of some inherent limitations. The through the simplification of the process flowsheet by
relative volatilities of the reactants and the products the elimination of some process units and direct heat
should be such that the products can be removed from integration between reaction and separation in the same
the reaction zone of the column easily while the reac- vessel.
On the other hand, because reactive distillation
* To whom correspondence should be addressed. Tel.: 610- systems have a smaller number of control degrees of
758-4256. Fax: 610-758-5057. E-mail: wll0@lehigh.edu. freedom than conventional multiunit reactor/separation/
10.1021/ie030832g CCC: $27.50 © 2004 American Chemical Society
Published on Web 04/17/2004
2494 Ind. Eng. Chem. Res., Vol. 43, No. 10, 2004

recycle systems, they can have worse dynamics, and conventional multiunit reactor/separator/recycle struc-
their controllability can be more difficult. For a given ture and (2) a reactive distillation column. Both of the
feed flow rate, a reactive distillation column has the flowsheets are designed to achieve the steady-state
same control degrees of freedom as a single conventional economic objective of minimum total annual cost for a
column (four control valves give four control degrees of wide range of chemical equilibrium constants, KEQ.
freedom). A conventional multiunit flowsheet has four Systematic design procedures for both flowsheets are
control degrees of freedom for each column, plus the two developed as three-dimensional optimization problems
control valves on the reactor (coolant flow rate and with some heuristic rules and engineering assumptions.
reactor effluent). The effects of design parameters on the optimum design
One of the most fundamental differences between variables and total annual costs are studied for both
reactive distillation and a conventional flowsheet is the process designs. The two flowsheets have identical feeds
selection of operating temperatures. In the conventional and produce identical products.
system, the reactor temperature can be set at an The vapor-liquid equilibrium used in this paper is
optimum value, and the distillation temperatures can ideal with constant relative volatilities. The rationale
be independently set at their optimum values (by for considering ideal systems is to start with a simple
adjusting column pressures). In reactive distillation, system in which the complexities of nonideality do not
these temperatures are not independent. Therefore, cloud the issues and the results and some general
reactive distillation has fewer design degrees of freedom conclusions are possible. The occurrence of azeotropes
to be used to optimize the flowsheet. is very specific to the particular chemical system
The application of reactive distillation in the chemical considered. We wish to strip away such complexities so
and petroleum industries has increased rapidly in the that the fundamental differences between the two
past decade.2 Following the pioneering paper3 published flowsheets can be fairly compared. Future work will
by Eastman Chemical discussing the reactive distilla- explore the impact of temperature-dependent relative
tion of methyl acetate, many other papers and patents volatilities on the two processes. It is expected that there
have explored the use of this process for other chemical will be a significant effect on the comparison because
operations. A number of specific chemical systems have of the coupling of the distillation temperatures and the
been studied in the literature. Chapter 10 of the recent reaction temperatures in the reactive distillation pro-
textbook by Doherty and Malone4 presents an excellent cess. In the conventional flowsheet, the reactor temper-
summary of 61 different reactive distillation systems. ature and the column temperatures can be set indepen-
Another list of industrial applications of reactive distil- dently.
lation with 75 examples is given in the first chapter of
a recent book on reactive distillation.5 This work states 2. Process Studied
that the most common applications of reactive distilla- The basic process considered consists of a reversible
tions are etherification and esterification reactions. liquid-phase reaction
Although reactive distillation can be effectively used for
systems with a wide variety of chemical equilibrium A+BSC+D (1)
constants, the normal range of values for KEQ is between
1 and 50. The forward and backward specific reaction rates,
Because a reactive distillation column has more following the Arrhenius law, are given by
design variables than a conventional column, the design
of such systems is more difficult. The column pressure, kF ) aFe-EF/RT (2)
the catalyst holdup on each tray, and the positions of
the feed trays are important design considerations. In kR ) aRe-ER/RT (3)
addition to stripping and rectifying sections, there is one
more column section (the reactive zone). Also, the The rate law is based on concentrations in mole frac-
assumption of constant molar overflow is not valid if tions and liquid holdups in kilomoles. The forward
the reactions are not thermally neutral. Several papers reaction rate is specified as 0.008 kmol s-1 kmol-1 at
have optimized reactive distillation columns by mini- 366 K. Taking (KEQ)366 equal to 2 as a base case, the
mizing the objective function in terms of total annual reverse reaction rate at this temperature is varied by
cost (TAC).6-10 All of these papers proposed different selecting a range of (KEQ)366 values between 0.5 and 50.
optimization methods for reactive distillation systems
and illustrated their methods with examples of specific (kF)366
chemical systems. (kR)366 ) (4)
(KEQ)366
Only a few papers have appeared in the literature
that compare reactive distillation flowsheets with other Therefore, for different values of (KEQ)366, different
process flowsheets.11-13 Siirola11 claimed that both the values of the preexponential factor aR are calculated.
capital and energy costs of the reactive/extractive distil- Both reaction rates are temperature-dependent, and the
lation column design for methyl acetate production are reverse reaction rate is different for each case of (KEQ)366
5 times lower than those of the conventional design. The selected. Note that the ratio of kF to kR is not equal to
results in terms of TAC reported by Chiang et al.12 for (KEQ)366 at temperatures other than 366 K because the
the amyl acetate process indicated that reactive distil- activation energies are different. Rather, because the
lation is 4 times more efficient than the coupled reactor/ reaction is exothermic, the reverse reaction rate is more
separator system. However, Stitt13 claimed that reactive temperature-dependent than the forward reaction rate.
distillation for toluene disproportionation did not offer Ideal vapor-liquid equilibrium is assumed with con-
significant economic benefits. stant volatilities. The volatilities of the components are
The purpose of this paper is to quantitatively compare such that RC > RA > RB > RD. Hence, the reactants are
the designs of two different process flowsheets: (1) a intermediate boiling between the two products, which
Ind. Eng. Chem. Res., Vol. 43, No. 10, 2004 2495

Table 1. Physical Data for the Process 3. Conventional Process Design


parameter units value
Figure 1 presents a detailed flowsheet of the process
reaction activation energy cal/mol with notation. The reaction occurs in a continuous
forward 30 000 stirred tank reactor (CSTR) with holdup VR. There are
reverse 40 000
specific reaction rate at 366 K kmol/(s‚kmol)
two fresh feed streams F0A and F0B that contain pure
forward 0.008 reactants A and B, respectively, and a recycle stream
reverse 0.008/(KEQ)366 D2 that returns from a downstream unit. The reactor
average heat of reaction, λ cal/mol -10 000 effluent contains a multicomponent mixture because
average heat of vaporization, cal/mol 6944 complete one-pass conversion is not achieved. Two
∆HV columns are needed to separate the two products from
molecular weight of the g/mol 50
mixture, Mw the intermediate-boiling reactants. This effluent is fed
ideal gas constant cal/(mol K) 1.987 into the first distillation column to separate product C
relative volatilities from unreacted reactants A and B and heavy product
RA 4 D. The product C leaves in the distillate of first column
RB 2 with the desired purity, and other components serve as
RC 8 a feed to the second column. This column produces a
RD 1
vapor-pressure A B C D
bottoms stream of D with the desired purity, and a
constantsa distillate of unreacted reactants A and B is recycled back
AVP 12.34 11.65 13.04 10.96 to the reactor with a specific amount of product impuri-
BVP 3862 3862 3862 3862 ties.
a ln PSj ) AVP,j - BVP,j/T, with temperature in Kelvin and vapor One of the main problems in developing a steady-state
pressure in bar. process flowsheet is finding the number of design
degrees of freedom, which can be done by subtracting
the number of chemical and physical equations (total
is the ideal situation for reactive distillation. Kinetic and
and component balances, VLE relations) describing the
physical properties and vapor-liquid equilibrium pa- system from the total number of variables (see Table
rameters are listed in Table 1 and are taken from 2). The number of design degrees of freedom for this
Luyben.14 multiunit process is 12. Subtracting the number of
It should be emphasized that the two flowsheets have specifications and safety and environmental constraints
identical feeds and produce identical products. In ad- from the number of degrees of freedom gives the number
dition, the economic factors (energy cost and capital cost of design optimization variables.
of columns and heat exchangers) are the same in the 3.1. Assumptions and Specifications. It is as-
two processes. sumed that there is equimolal overflow in the distilla-

Figure 1. Detailed system flowsheet for the conventional design.


2496 Ind. Eng. Chem. Res., Vol. 43, No. 10, 2004

Table 2. Design Degrees of Freedom


variables equations
compositions component balances
reactors zj 3 reactor 3
tray liquid xn,j 3∑N trays 3∑N
tray vapor yn,j 3∑N bases 6
reflux drums xD,j 6 reflux drums 6
base liquid xB,j 6 total balances
base vapor yB,j 6 reactor 1
flows bases 2
fresh feeds F0,j 2 reflux drums 2
reactor effluent F 1 VLE
vapor boilups Vn 2 trays 3∑N
distillates Dn 2 bases 6
bottoms Bn 2
refluxes Rn 2
reactor holdup VR 1
reactor temperature TR 1
numbers of trays
stripping NS,n 2
rectifying NR,n 2
total number of variables ) 6∑NT + 38 total number of equations ) 6∑NT + 26
number of design degrees of freedom ) 12

tion columns, which means that neither energy balances search with the following optimization variables: (1) the
nor total balances are needed on the trays for steady- molar holdup in the reactor VR, (2) the composition of
state calculations. Other assumptions are constant reactant B in the reactor zB, and (3) the reactor
relative volatilities, isothermal reactor operation, theo- temperature TR.
retical trays, saturated liquid feed and reflux, and total Also, several different specifications for assumptions
condensers and partial reboilers in the columns. Ad- ii-v are used to investigate the effects of product
ditional assumptions and specifications are as follows: quality, conversion and recycle impurities on the eco-
(i) In all cases, the net production rates are set by nomically optimal steady-state design. However, these
fixing the fresh feed flow rates of pure components A variations do not affect the general structure of the
and B at F0A ) 12.60 mol/s and F0B ) 12.60 mol/s, design procedure.
respectively. A grid-search optimization strategy is used in this
(ii) The amount of reactant A lost in product stream work to find the optimum values of the three design
D1 is constant at Aloss ) 0.63 mol/s (xD1,A ) 0.05 mole optimization variables. Other methods, such as gradient
fraction). No B or D is going overhead, i.e., xD1,B ) xD1,D nonlinear programming techniques, could be used.
) 0. However, the grid method is more robust because a
(iii) The bottoms product stream B2 contains some numerical programming method can easily drive the
amount of component B as an impurity, Bloss ) 0.63 process into an infeasible region in which the specified
mol/s (xB2,B ) 0.05 mole fraction). No A or C is going purities and production rates cannot be achieved.
out the bottoms, i.e., xB2,A ) xB2,C ) 0. 3.2. Steady-State Design Procedure. The following
(iv) The impurity of product C in recycle stream D2 is steps in the design procedure are employed utilizing the
xD2,C ) 0.05. material balances and specifying necessary variables.
(v) The impurity of component D in recycle stream The production rates of components C and D (RC and
D2 is xD2,D ) 0.05. RD) are given by
(vi) As a result of the previous two assumptions and
the fact that C and D have identical stoichiometric RC ) RD ) VR(kFzAzB - kRzCzD) (5)
coefficients, the concentrations of components C and D
in the reactor, zC and zD, are equal for all cases (zC ) (i) Fix the value of reactor temperature TR at a small
zD) but vary in value from case to case. value.
(vii) The column pressures are set using the vapor (ii) Fix the value of reactor holdup VR at a small value.
pressures, PS, of pure components and the liquid (iii) Specify the value of reactor composition zB.
compositions in the reflux drum, xD,j, at 320 K (so that (iv) Using the assumption of equal compositions zC
cooling water can be used in the condenser). and zD in the reactor, calculate the concentration zA by
(viii) In each column, the reflux ratio, RR, is 1.2 times rearranging eq 5 to give

{
the minimum reflux ratio, RRmin, calculated via the
Underwood equations.
1 kF
(ix) In each column, the number of stages (or trays), zA ) 4 z + 2(1 - zB) (
2 kR B

]}
NT, is twice the minimum number of stages, NTmin,

x[ ] [
calculated via the Fenske equation. kF 2 RC
(x) Kirkbride’s method is used to find the optimal feed 4 zB + 2(1 - zB) - 4 (1 - zB)2 + 4 (6)
tray location, NF.
kR VRkR
Through these assumptions, nine of the design de-
grees of freedom are used as follows: one for the (v) Calculate the product concentrations zC and zD in
production rate, two for product impurities, two for the reactor as
recycle impurities, and four for total tray numbers and
optimum feed tray locations. This reduces the steady- 1 - zA - zB
zC ) z D ) (7)
state economic optimization to a three-dimensional 2
Ind. Eng. Chem. Res., Vol. 43, No. 10, 2004 2497

(vi) Calculate the distillate product flow rate and


compositions for the first column

D1 ) RC + Aloss (8)

xD1,A ) Aloss/D1 (9)

xD1,C ) RC/D1 (10)

(vii) Calculate the bottoms flow rate and compositions


for the second column

B2 ) RD + Bloss (11)

xB2,B ) Bloss/B2 (12)

xB2,D ) RD/B2 (13)

(viii) Apply a steady-state total molar balance, which


can be used because the reactions are equimolar. This
balance, the two steady-state component balances, and
a mole fraction summation around the reactor are

F0A + F0B + D2 ) F (14)

F0A + D2xD2,A ) FzA + RC (15)

F0B + D2xD2,B ) FzB + RD (16)

xD2,A + xD2,B + xD2,C + xD2,D ) 1 (17)


Figure 2. Reactive distillation column.
Solve these four equations for the four unknowns: the
recycle stream flow rate D2, the reactor effluent flow (xi) Using these data, calculate the total annual cost
rate F, and the reactant compositions in second distillate (TAC) by combining the energy cost with the annual
stream xD2,A and xD2,B. The combination of eqs 14-17 capital cost, using a payback period. All of the terms
yields and equations related to the sizing and economy of the
process are given in section 5.
(F0A + F0B)(zA + zB - 1) + RC + RD (xii) Vary the value of zB over a wide range, and repeat
D2 ) (18) steps iv-xi for each value of zB, generating the corre-
(1 - xD2,C - xD2,D) - zA - zB
sponding TAC.
(xiii) Then, vary the value of the reactor holdup over
F ) F0A + F0B + D2 (19) a range, and repeat steps iii-xii.
(xiv) Finally, vary the value of the reactor tempera-
FzA + RC - F0A ture over a wide range, and repeat steps ii-xiii for each
xD2,A ) (20)
D2 temperature. Select the minimum in the TAC as the
economically optimum steady-state design for the given
xD2,B ) (1 - xD2,C - xD2,D) - xD2,A (21) (KEQ)366 value.

(ix) Use the total mass balance and component 4. Reactive Distillation Design
balances around the first column to calculate the
The reactive distillation alternative is shown in
bottoms flow rate and compositions for that column
Figure 2. The column is fed with two pure reactant fresh
B1 ) F - D1 (22) feed streams: F0A and F0B. The column has three zones.
There are NS trays and a partial reboiler in the stripping
Fzj - D1xD1,j section. Above this section, there is a reactive zone with
xB1,j ) (23) NRX reactive trays. The third section is the rectifying
B1 section with NR trays and a total condenser.
Light reactant A is fed to the bottom tray of the
(x) Now that the feed, bottoms, and distillate flow reactive zone, while heavy reactant B is introduced at
rates and compositions are known for both columns, use the top of the reactive section. Light product C leaves
the Fenske equation to calculate the minimum number in the distillate, while heavy product D is removed from
of trays. Set the actual number of trays equal to twice the bottoms. Because the light reactant goes up through
the minimum. Then use the Underwood equations to the column after being fed on the bottom tray of the
calculate the minimum reflux ratio. For sizing purposes, reactive zone, very little of component A is present in
set the actual reflux ratio equal to 1.2 times the the bottoms. Likewise, heavy reactant goes down through
minimum value. the column after being fed on the top tray of the reactive
2498 Ind. Eng. Chem. Res., Vol. 43, No. 10, 2004

zone, so very little of component B is present in the


distillate. Thus, the key components are light product
C and light reactant A for the rectifying section. For
the stripping section, the key components are heavy
product D and heavy reactant B.
4.1. Assumptions and Specifications. The optimi-
zation of a reactive distillation column has a very large
number of design variables. The following are the
specifications and assumptions made to reduce the
number of design variables for the economically opti-
mum steady-state design:
(i) The design objective is to obtain 95% conversion
for fixed fresh feed flow rates of 12.6 mol/s.
(ii) The product impurity of the distillate stream is
xD,A ) 0.05.
(iii) The product impurity of the bottoms stream is
xB,B ) 0.05.
(iv) The two feed points are at the two ends of the
reactive zone.
(v) The holdups are constant on the reactive trays at
1000 mol and in the other sections of the column at 400
mol.
(vi) The numbers of stripping and rectifying trays are
equal because the relative volatilities between the
components being separated in the two sections are the Figure 3. Reactive tray.
same. In the rectifying section, the separation is be-
tween C and A, where the relative volatility is RC/RA )
2, and in the stripping section, the separation is between the liquid flow rate decreases down through the reactive
B and D, where the relative volatility is RB/RD ) 2. zone.
With these specifications and simplifying assump-
λ
tions, there are three optimization variables: (1) the Vi ) Vi-1 - R (25)
column pressure P, (2) the number of reactive trays NRX, ∆HV i,C
and (3) the number of the separating (stripping/rectify-
ing) trays NS (or NR). λ
Li ) Li+1 + R (26)
Specifications ii and iii are changed during the study ∆HV i,C
to investigate the effects of product quality and conver-
sion on the economically optimal steady-state design. The dynamic component balances for the column are
However, this does not affect the general structure of as follows:
the design procedure.
4.2. Steady-State Design Procedure. Simulta- reflux drum
neous solution of the very large set of nonlinear and d(xD,jMD)
algebraic equations is difficult, especially with the high ) VNTyNT,j - D(1 + RR)xD,j (27)
degree of nonlinearity due to reaction kinetics. The dt
relaxation method is efficient and robust in solving this
large set of equations. This method is used to calculate rectifying and stripping trays
mole fractions and temperature profiles through the d(xi,jMi)
column. In general, relaxation methods use the equi- ) Li+1xi+1,j + Vi-1yi-1,j - Lixi,j - Viyi,j (28)
librium-stage model equations in unsteady-state form dt
and integrate them numerically until the steady-state
reactive trays
solution is found. Here, the liquid holdups on the trays
are assumed constant, i.e., instantaneous hydraulics. A d(xi,jMi)
general schematic diagram of an equilibrium tray in the ) Li+1xi+1,j + Vi-1yi-1,j - Lixi,j - Viyi,j + Ri,j
dt
reactive zone, from which the equations for a dynamic (29)
model can be derived, is shown in Figure 3. The net
reaction rate for component j on tray i in the reactive feed trays
zone is given by
d(xi,jMi)
) Li+1xi+1,j + Vi-1yi-1,j - Lixi,j - Viyi,j +
Ri,j ) νjMi(kFixi,Axi,B - kRixi,Cxi,D) (24) dt
Ri,j + Fizi,j (30)
where νj is the stoichiometric coefficient of component
j. column base
The steady-state vapor and liquid flow rates are
constant through the stripping and rectifying sections d(xB,jMB)
because equimolal overflow is assumed. However, these ) L1x1,j - BxB,j - VSyB,j (31)
dt
rates change through the reactive zone because of the
exothermic reaction. The heat of reaction vaporizes some With the equimolal overflow assumption mentioned
liquid on each tray in that section; therefore, the vapor above, all of the vapor rates, Vi, throughout the stripping
flow rate increases up through the reactive zone, and section are equal to VS, and all of the liquid rates, Li,
Ind. Eng. Chem. Res., Vol. 43, No. 10, 2004 2499

are equal to LS. Analogously, all of the vapor rates, Vi, Table 3. Sizing and Economic Basis
beginning from the top feed tray throughout the rectify- parameter units value
ing section and total condenser are VNT, and all of the
reboiler
liquid rates, Li, are equal to LR. heat-transfer coefficient, UR kJ/(s‚K‚m2) 0.568
The vapor-liquid equilibrium is assumed to be ideal. temperature difference, ∆TR K 34.8
Column pressure P is optimized for each case. With condenser
given pressure P and tray liquid composition xi,j, the heat-transfer coefficient, UC kJ/(s.K.m2) 0.852
temperature Ti, and the vapor composition yi,j can be temperature difference, ∆TC K 13.9
calculated. This is a bubble-point calculation and can energy cost $/106 kJ 4.7
payback period, βpay year 3.0
be solved by a Newton-Raphson iterative convergence
method. can be easily observed from the continuously increasing
NC vapor boilup, and the simulation can be restarted in
P) ∑
j)1
S
xi,jPi,j(T) (32)
such cases with new optimization variables.

5. Sizing and Economic Basis


S
Pi,j The economic basis for the calculations involves an
yi,j ) x (33) objective function that sums the capital and energy costs
P i,j
of the process assuming a payback period (βpay) for
With a fixed value of the kinetic parameter (KEQ)366, capital. Total annual cost is defined as
the following steps in the design procedure are used:
(i) Fix the column pressure P at a small value. capital investment
TAC ) energy cost + (34)
(ii) Fix the number of stripping trays NS and the βpay
number of rectifying trays NR at a small value.
(iii) Specify the number of the reactive trays NRX. Table 3 summarizes the economic parameters and the
(iv) Fix the flow rates of the two fresh feeds at 12.60 basis of the equipment sizing calculations.
mol/s. The capital costs of individual equipment and the
(v) Fix the flow rates of the distillate and bottoms at energy cost are estimated using the following equa-
12.60 mol/s. tions15
(vi) Manipulate the vapor boilup VS with a P control-
ler to control the level in the column base. Do not control reactor cost ) 52 920DR1.066LR0.802 (35)
the reflux drum level.
(vii) Manipulate the reflux flow rate with a PI column cost ) 17 640DC1.066LC0.802 (36)
controller to drive the composition of product C in the
distillate to its desired value. This also sets the purity
of the bottoms product D and the conversion in the tray cost ) 229DC1.55NT (37)
reactive zone to their desired values.
(viii) Compute the vapor compositions and tempera- heat exchanger cost ) 7296AR0.65 + 7296AC0.65 (38)
tures on each tray using bubble-point calculations.
(ix) Compute the reaction rates with eq 24. energy cost ) 0.6206∆HVVS (39)
(x) Calculate the vapor rates using eq 25 and assum-
ing equimolal overflow through the stripping and rec- The terms involved in the TAC equations are calcu-
tifying sections. lated from the following set of assumptions and guide-
(xi) Calculate the liquid flow rates from eq 26 with lines:
the same assumption as in step x. (i) The diameter of the reactor is calculated from
(xii) Evaluate the time derivatives of the component
material balances using eqs 27-31. DR ) (3.977 × 10-5VR)0.3333 (40)
(xiii) Integrate all of the ODEs using the Euler
algorithm. (ii) The reactor length is assumed to be twice the
(xiv) Repeat steps vi-xiii until the system achieves diameter.
the convergence criterion, which is CC ) max |dxi,j/dt|
e 10-6 (the largest time derivative of any component LR ) 2DR (41)
on any tray is less than 10-6).
(xv) With the data obtained, calculate the total annual (iii) Assuming an F factor of 1 in engineering units,
cost (TAC) by combining the energy cost with the annual the diameter of the column is calculated from the
capital cost. equation
(xvi) Vary the number of reactive trays over a wide
range, and repeat steps iv-xv for each value of NRX,
generating its corresponding TAC.
(xvii) Then, vary the number of the stripping and
DC ) 1.735 × 10-2 ( )
MwT
P
0.25
VNT0.5 (42)

rectifying trays over a range, and repeat steps iii-xvi. (iv) The column height is calculated assuming a
(xviii) Finally, vary the value of the column pressure 0.61-m (2-ft) tray spacing and allowing 20% more height
over a wide range, and repeat steps ii-xvii for each for base-level volume.
pressure value. Select the design with the minimum
TAC as the economically optimum steady-state design LC ) 0.731 52NT (43)
for the given (KEQ)366 value.
Depending on the values of optimization variables (v) The heat-transfer areas of the reboiler and con-
chosen, the column can leave the feasible region. This denser are calculated using the steady-state vapor flow
2500 Ind. Eng. Chem. Res., Vol. 43, No. 10, 2004

a minimum reactor effluent flow rate, and this occurs


in the region where the compositions of the two reac-
tants are similar because the reaction rate depends on
the product of the two reactant concentrations. The
optimum design point is shown on each graph as an
open circle.
Figure 6 shows that a larger reactor size results in a
higher capital cost but a lower energy cost because of
the lower vapor boilup requirement for all values of zB.
From the same figure, it is clear that for all reactor
holdups there is a minimum in both the capital cost and
energy cost at a specific value of zB at which the
concentrations of zA and zB are similar. Thus, there
should be a minimum in the total cost around this value
of zB because of the tradeoff between increasing capital
Figure 4. TAC for the base case of (KEQ)366 ) 2.00 with constant cost and decreasing energy cost.
reactor temperature TR ) 367 K.
Figures 7 and 8 provide results for the base case using
a range of reactor temperatures as the third optimiza-
rates and the heat of vaporization.
tion variable. All of the results in both figures are the
economically optimal steady-state values at the given
VS∆HV temperature. The specific reaction rates in the reactor
AR ) 0.0042 (44)
UR∆TR increase with increasing temperatures. Figure 7 shows
that increasing the temperature decreases the reactor
VNT∆HV holdup (VR) for a specified rate of production of C and
AC ) 0.0042 (45) D. This means that higher temperature results in lower
UC∆TC
reactor cost, as shown in Figure 8. On the other hand,
the column cost, the heat exchanger cost, and the energy
The vapor flow rate in the top tray, VNT, is higher than cost all depend on the vapor boilups in the system. It
the vapor flow rate in the reboiler, VS, because of the can be seen from Figure 7 that there are minima in the
liquid vaporized through the reactive section. Thus, the vapor boilup curves at certain values of reactor tem-
heat-transfer areas of the reboiler and condenser are perature. Because the optimum value of VR decreases
calculated using two different vapor rates. with increasing reactor temperature, the per-pass con-
(vi) The two processes are assumed to be equally version in the reactor decreases. Thus, the reactor
reliable and to operate for 365 days per year. effluent F contains more reactant components A and B
at higher temperatures. Another reason for the increase
6. Results and Discussion in the reactant contents in the reactor effluent is that
6.1. Conventional Process Design. Figure 4 shows the equilibrium constant KEQ decreases with increasing
a plot of the total annual cost (TAC) for the base case, temperature for exothermic reactions. More reactants
(KEQ)366 ) 2, with the constant reactor temperature TR fed to the columns result in an increase in reflux and
) 367 K. The TAC is plotted versus the reactor holdup, vapor boilup required to achieve the desired purities of
VR, and the composition of component B in the reactor, product streams. Therefore, these separation costs
zB. The reactor holdup and the reactor composition of increase when the reactor temperature gets high.
component B were varied in increments of 2.5 kmol and Because of the decreasing reactor cost, the total capital
0.025, respectively, yielding a total of 360 different cost decreases with the increase in temperature up to a
designs. certain value. Figure 8 shows, however, that above this
With an in-depth look, Figure 5 displays how several temperature, the increasing column cost and heat
design parameters are affected by the variation of the exchanger cost are more dominant and result in in-
optimization variables VR and zB, and Figure 6 shows creasing total capital cost.
how capital cost and energy cost vary. Both figures are The total annual cost reaches a minimum at a certain
plotted versus zB for three different values of reactor reactor temperature because of the tradeoff between
holdup VR with the fixed temperature TR ) 367 K. reactor cost and separation cost. The optimum reactor
Figure 5 demonstrates that the reactor effluent temperature for this base-case value of (KEQ)366 is 367
contains a higher concentration of reactant components K. The optimum values of the other optimization
as the reactor size decreases. Thus, there is a need of variables at the optimum temperature are VR ) 102.5
more vapor boilup to achieve the desired product kmol and zB ) 0.225. Other important design param-
specifications and an increase in recycle flow rate with eters are the recycle flow rate D2 (22.89 mol/s) and the
smaller reactor holdup. Increasing zB decreases zA for composition of component A in the reactor (zA ) 0.2296).
a given reactor holdup. The vapor boilup in the first Figure 9 presents results for different (KEQ)366 values.
column, VS1, decreases because there is less A to be The first graph shows that at higher values of (KEQ)366
separated from C. As the value of zB increases, the vapor the total annual cost (TAC) of the system decreases. The
boilup in the second column, VS2, starts to increase optimum temperature increases as the value of (KEQ)366
because there is more B to take overhead. The process increases. Reactor holdups decrease for all kinetic cases
can be operated at any point on this curve with the same as the temperatures increase. The lower graphs show
production rate for this given VR, but the reactor effluent that higher values of (KEQ)366 result in lower values of
F, the flow rate of bottoms stream B1, and the flow rate the recycle stream and vapor boilup. The curves of the
of recycle stream D2 will vary, as shown in the lower recycle stream and vapor boilup of second column
graphs of Figure 5. There is a value of zB that produces exhibit minima at certain values of reactor temperature.
Ind. Eng. Chem. Res., Vol. 43, No. 10, 2004 2501

Figure 5. Effects of optimization variables VR and zB for the base case of (KEQ)366 ) 2.00 with the constant temperature TR ) 367 K.

Figure 6. Effects of optimization variables VR and zB on design Figure 7. Effects of optimization variable TR for the base-case
costs for the base case of (KEQ)366 ) 2.00 with the constant (KEQ)366 ) 2.00 design.
temperature TR ) 367 K.
sures are set using the vapor pressures, PS, of the pure
For each kinetic case, the shapes of the curves are components at 320 K and the liquid compositions in the
similar, but they have minima at different tempera- reflux drum, xD,j. In a conventional multiunit process,
tures. For both low and high temperatures, the increase the column temperatures can be set independently so
of these flow rates occurs because of the higher reactant as to optimize column efficiency. The reactor tempera-
compositions in the reactor effluent, as shown in Figure ture can then be independently adjusted to its optimum.
7 for the base case, (KEQ)366 ) 2. As mentioned before, This is not the case in reactive distillation because both
higher temperatures give smaller reactor holdups VR reaction and separation are occurring in the same
and lower equilibrium constants KEQ, which give higher vessel, operating at a single pressure.
reactant concentrations. However, low temperatures
give small specific reaction rates, meaning that more The effect of the conversion design value is explored
reactant leaves the reactor. Therefore, higher vapor by increasing the conversion from 95 to 99%. Increasing
boilup is required in the columns to obtain the product the conversion means that the product purities increase
streams with specified purities. (less unreacted A and B are lost in the product streams)
Optimization results for the conventional design for and the optimum values of all of the design parameters
all cases are summarized in Table 4. These are the change. Figure 10 gives results for all of the (KEQ)366
optimum designs in terms of the reactor temperature, cases. The vapor boilup required to achieve 99% conver-
the reactor holdup, and the composition of component sion increases, as does the flow rate of the recycle
B in the reactor. Optimal operating conditions for the stream, D2. The reactor holdups are almost the same
same cases are reported in Table 5. The column pres- for both conversions. These results illustrate that, as
2502 Ind. Eng. Chem. Res., Vol. 43, No. 10, 2004

Figure 8. Effects of optimization variable TR on design costs for the base-case (KEQ)366 ) 2.00 design.

C and D are simultaneously varied from 0.001 to 0.125.


The TAC decreases with higher amounts of impurities.
The reactor holdup also decreases, whereas the recycle
flow rate and vapor boilup of the second column
increase. Although there is an initial large decrease in
TAC as impurities are increased from 0.001, the change
in TAC is not very significant above an impurity value
of about 0.05. Therefore, xD2,C ) xD2,D ) 0.05 is used as
the base impurity level for other kinetic cases.
6.2. Reactive Distillation Design. Figure 12 shows
the effects of separation stages on the economically
optimum steady-state design of a reactive distillation
column with three different operation pressures. The
results are given for the base case, (KEQ)366 ) 2, with a
number of constant reactive trays, NRX ) 9. The graph
in the middle shows that the column cost increases as
the number of separation trays increase. It also shows
Figure 9. Optimization results for different kinetic cases. that a decrease in the operation pressure results in an
increase in the column cost for any number of separation
Table 4. Optimization Results for the Conventional
Design
trays because decreasing the pressure results in a lower
density, which increases the column diameter. The right
(KEQ)366 graph indicates that the higher the number of separa-
0.5 1.0 2.0 5.0 10.0 50.0 tion stages, the lower the vapor boilup required. More
design variables stripping and rectifying stages provide the required
TR (K) 356.0 362.0 367.0 373.0 379.0 395.0 separation while using less energy. Because the tem-
VR (kmol) 222.5 145.0 102.5 82.5 60.0 25.0 perature of the reactive zone increases with increasing
zB 0.275 0.250 0.225 0.175 0.150 0.110 column pressure, it decreases the chemical equilibrium
capital costs ($103)
reactor 358.6 274.7 221.3 193.3 158.6 91.9
constant, which results in an increase in the amounts
heat exchanger 630.1 565.7 509.6 444.2 407.2 341.3 of reactant leaving the reactive zone. Therefore, the
column 293.3 268.1 245.8 217.9 202.4 175.0 right graph also shows that operating at higher pressure
tray 9.5 8.3 7.3 6.1 5.5 4.4 increases the vapor boilup required to meet desired
energy cost ($103/year) 502.9 425.4 361.9 292.8 256.3 195.7 product specifications. The left graph of Figure 12 shows
TAC ($103/year) 933.4 797.7 689.9 580.0 514.2 399.9
that there is a minimum in the total annual cost curve
at a certain number of separation trays because of the
expected, the TAC of the system increases when a tradeoff between increasing column cost and decreasing
higher conversion is required. costs related to the vapor boilup, such as energy and
Figure 11 shows the effect of impurities in the D2 heat exchanger costs. The overlapping TAC curves for
recycle stream on the total annual cost for the base case, different pressures show that there is an optimum
where (KEQ)366 is 2. The impurities of both components pressure. This occurs because of the competing effects
Ind. Eng. Chem. Res., Vol. 43, No. 10, 2004 2503

Table 5. Optimal Operating Conditions of Conventional


Design
(KEQ)366
parameter 0.5 1.0 2.0 5.0 10.0 50.0
Reactor
F (mol/s) 66.02 55.81 48.09 40.39 36.65 31.20
VR (kmol) 222.50 145.00 102.50 82.50 60.00 25.00
TR (K) 356.00 362.00 367.00 373.00 379.00 395.00
kF (s-1) 0.0025 0.0051 0.0090 0.0174 0.0329 0.1653
kR (s-1) 0.0034 0.0044 0.0046 0.0045 0.0053 0.0091
zA 0.3005 0.2662 0.2296 0.1947 0.1655 0.1035
zB 0.2750 0.2500 0.2250 0.1750 0.1500 0.1100
zC 0.2122 0.2419 0.2727 0.3151 0.3422 0.3933
zD 0.2122 0.2419 0.2727 0.3151 0.3422 0.3933
Column 1
B (mol/s) 53.42 43.21 35.49 27.79 24.05 18.60
D (mol/s) 12.60 12.60 12.60 12.60 12.60 12.60
VS (mol/s) 50.56 44.57 39.56 34.67 31.66 25.62
R (mol/s) 37.96 31.97 26.96 22.07 19.06 13.02
P (bar) 2.57 2.57 2.57 2.57 2.57 2.57 Figure 10. Effects of conversion for different kinetic cases.
xD,A 0.0500 0.0500 0.0500 0.0500 0.0500 0.0500
xD,B 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000
xD,C 0.9500 0.9500 0.9500 0.9500 0.9500 0.9500
xD,D 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000
xB,A 0.3596 0.3292 0.2933 0.2603 0.2261 0.1397
xB,B 0.3399 0.3229 0.3049 0.2543 0.2286 0.1845
xB,C 0.0382 0.0354 0.0322 0.0273 0.0238 0.0161
xB,D 0.2623 0.3124 0.3695 0.458 0.5215 0.6596
NT 13 13 13 13 13 13
NF 6 6 7 7 8 8
DC (m) 1.09 1.03 0.97 0.90 0.86 0.78
AR (m2) 74.36 65.56 58.18 50.99 46.57 37.68
AC (m2) 124.11 109.42 97.11 85.10 77.72 62.90
Column 2
B (mol/s) 12.60 12.60 12.60 12.60 12.60 12.60
D (mol/s) 40.82 30.61 22.89 15.19 11.45 6.00
VS (mol/s) 66.14 54.15 44.42 33.28 27.82 19.80
R (mol/s) 25.33 23.54 21.54 18.09 16.37 13.80
P (bar) 1.00 0.99 0.98 1.00 1.00 0.97
xD,A 0.4706 0.4648 0.4548 0.4762 0.4749 0.4331
xD,B 0.4294 0.4352 0.4452 0.4238 0.4251 0.4669
xD,C 0.0500 0.0500 0.0500 0.0500 0.0500 0.0500
xD,D 0.0500 0.0500 0.0500 0.0500 0.0500 0.0500 Figure 11. Effects of impurities in the recycle stream for the base
xB,A 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 case of (KEQ)366 ) 2.0.
xB,B 0.0500 0.0500 0.0500 0.0500 0.0500 0.0500
xB,C 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000
xB,D 0.9500 0.9500 0.9500 0.9500 0.9500 0.9500
NT 13 13 13 13 13 13
NF 8 8 7 7 6 6
DC (m) 1.58 1.43 1.30 1.12 1.03 0.87
AR (m2) 97.29 79.65 65.34 48.96 40.92 29.12
AC (m2) 162.38 132.93 109.06 81.71 68.30 48.61

of temperature on the reaction rates and the chemical


equilibrium constant. For this case, five stripping and
five rectifying trays are optimal for all pressures.
Figure 13 reveals the effects of the number of reactive
trays for the base case with the constant number of
separation stages NR ) NS ) 5. The graph in the middle
shows that increasing the number of reactive trays
results in an increase in the column cost and that higher
pressures lead to a lower column cost for any value of
NRX.
The right graph shows the effects of the number of
reactive stages on the vapor boilup. As the number of Figure 12. Effect of number of separation stages for the base
reactive trays increases, the required vapor boilup case of (KEQ)366 ) 2 with constant NRX ) 9.
initially decreases. However, increasing the number of
reactive trays above the optimum results in an increase cussed by Sneesby and co-workers.16 For example, the
in the required vapor boilup. This occurs because the concentration of reactant A on the lowest reactive tray
extra separation occurring with more stages concen- increases from 0.4425 to 0.5179 as NRX is changed from
trates the products in the reactive zone and shifts the 9 (the optimum) to 11. The concentration of product D
chemical equilibrium back toward the reactants on the at the same location decreases from 0.4817 to 0.4322.
lower and upper reactive stages. This effect was dis- Therefore, it takes more energy to keep component A
2504 Ind. Eng. Chem. Res., Vol. 43, No. 10, 2004

Figure 13. Effect of number of reactive stages for the base case Figure 15. Economically optimum results for three different
of (KEQ)366 ) 2 with constant number of separation stages NR ) kinetic cases.
NS ) 5.
optimum pressure is more important for small values
of (KEQ)366 than for large values. This figure shows that
the optimum pressures for (KEQ)366 ) 1, 2, and 5 are
6.5, 8.0, and 9.5 bar, respectively.
Figure 15 gives the economic optimum steady-state
design results for three different values of (KEQ)366. For
each case, the optimum pressure that minimizes the
vapor boilup is used, and the number of reactive trays
is varied over a range to find the minimum total annual
cost TAC. The upper graph shows that there is a
minimum in the TAC curve at some optimum number
of reactive stages. The optimum number of reactive
stages decreases with increasing values of (KEQ)366. The
column costs and total annual cost decrease as the value
of (KEQ)366 increases.
Results for five different kinetic cases are reported
in Table 6. The table includes the optimum design
parameters and costs in terms of the following optimi-
zation variables: (i) number of separation trays, (ii)
Figure 14. Effects of pressure on the vapor boilup for three
different kinetic cases.
number of reactive trays, and (iii) the column pressure.
Also, the temperatures at the top and bottom of different
from leaving in the bottoms. Thus, there is an optimum sections are included in this table.
number of reactive trays for each pressure that mini- Figure 16 shows the composition profile of the opti-
mizes the vapor boilup, energy cost, and heat exchanger mum design for the base case of (KEQ)366 ) 2. The
cost. highest composition of reactant A is at the bottom of
The left graph of Figure 13 shows that there is a reactive zone where A is fed. The other reactant, B, has
minimum in the total annual cost curve at a certain its highest composition at the top of reactive zone, which
number of reactive trays because of the tradeoff between is also its feed tray. While the composition of A
increasing column cost and decreasing vapor boilup. For decreases up through the reactive zone, the composition
this case, nine reactive trays are the optimum for all of product C increases. The reverse occurs for reactant
pressures. This figure also indicates that the optimum B and product D through the reactive zone. The rest of
pressure is 8 bar for this specific case. the column operates as a separation unit. Thus, the
Figure 14 shows the effect of operation pressure on composition of heavy product D increases down through
vapor boilup. These results are for the three different the stripping section, and the composition of light
kinetic cases (KEQ)366 ) 1, 2, and 5 with the number of product C increases up through the rectifying section.
reactive trays that is optimum for each case. The graph Figure 17 shows the temperature profiles of the
shows that there is an optimum pressure for all three optimum designs for the three different kinetic cases
cases that minimizes the vapor boilup. The higher (KEQ)366 ) 1, 2, and 5. The temperature profiles of the
values of (KEQ)366 require lower vapor boilups and three cases are similar, with higher temperatures for
operate at higher optimum column pressures. The higher values of (KEQ)366. Fairly significant temperature
higher equilibrium constant pushes the reaction to the breaks occur around tray 4 for all cases, and these tray
right. The equilibrium constant decreases with increas- temperatures can be used in control schemes to infer
ing temperature for an exothermic reaction. Therefore, bottoms purity. A similar break occurs at different tray
the column can operate at higher pressures when numbers near the top of the column for each case. These
(KEQ)366 is larger. It is possible to conclude from this temperatures could be used to infer distillate purity. For
figure that the sensitivity to pressure increases as the all kinetic cases, the temperatures show little change
value of (KEQ)366 decreases. Therefore, operating at the in the reactive zone.
Ind. Eng. Chem. Res., Vol. 43, No. 10, 2004 2505

Table 6. Optimization Results of Reactive Distillation Design


(KEQ)366
0.5 1.0 2.0 5.0 10.0 50.0
design variables
NR and NS 6 5 5 5 6 5
NRX 18 13 9 6 4 3
P (bar) 5.5 6.5 8.0 9.5 11.0 16.0
design temperatures (K)
base 415.8 423.1 432.8 441.0 449.0 467.6
bottom reactive 382.0 386.7 394.0 400.7 407.7 422.5
top reactive 365.1 383.3 393.9 402.6 409.7 428.4
reflux drum 341.5 346.6 353.2 358.9 363.8 377.3
design parameters
NT 30 23 19 16 16 13
VS (mol/s) 34.86 31.74 28.51 24.36 21.37 16.24
R (mol/s) 39.50 36.38 33.14 29.00 26.01 20.88
DC (m) 0.935 0.873 0.805 0.738 0.688 0.590
AR (m2) 51.28 46.68 41.93 35.83 31.44 23.89
AC (m2) 127.90 120.23 112.30 102.12 94.79 82.21
capital costs ($103)
heat exchanger 265.1 252.8 239.7 222.3 209.2 185.6
column 195.5 146.8 115.5 91.8 85.1 61.1
tray 6.2 4.3 3.1 2.3 2.1 1.3
energy cost ($103/year) 150.2 136.8 122.8 105.0 92.1 70.0
TAC ($103/year) 305.9 271.4 242.3 210.5 190.9 152.7

Figure 16. Composition profile of the optimum design for the Figure 18. Effect of conversion for the base case of (KEQ)366 )
base case of (KEQ)366 ) 2.0. 2.0.

at each pressure are given for base kinetic case of


(KEQ)366 ) 2. To achieve the required purity, the
optimum number of separation stages is NS ) NR ) 7
for this case, which is 2 more than in the 95% conversion
system. The optimum number of reactive stages is 13,
which is 4 more than in the base case. The optimum
pressure is the same (8 bar) as in the 95% conversion
case. These results indicate that the TAC of the system
increases when a higher conversion is required.
6.3. Comparisons. Comparisons of conventional
reactor/separator/recycle and reactive column design
configurations at their economically optimum steady-
state designs are presented in Table 7 for five different
kinetic cases.
The results indicate that the TACs of both design
configurations decrease as the value of (KEQ)366 in-
Figure 17. Temperature profiles of optimum designs for different
creases. The results also show that the reactive distil-
kinetic cases. lation column configuration has lower capital and
energy costs than the conventional configuration for all
Figure 18 illustrates the effect of the conversion on kinetic cases. These costs result in lower TACs for the
the total annual cost for different pressures. The reactive distillation columns compared to the reactor/
conversion is increased to 99%, and the optimal results column/recycle systems.
2506 Ind. Eng. Chem. Res., Vol. 43, No. 10, 2004

Table 7. Economic Comparison of Two Different Process Nomenclature


Flowsheets
aF ) preexponential factor for the forward reaction
(KEQ)366
(kmol‚s-1‚kmol-1)
0.5 1.0 2.0 5.0 10.0 50.0 aR ) preexponential factor for the reverse reaction
capital cost ($103) (kmol‚s-1‚kmol-1)
conventional 1291.5 1116.8 984.0 861.6 773.6 612.6 AC ) heat exchanger area for the condenser (m2)
reactive column 466.8 403.9 358.4 316.4 296.4 248.1 AR ) heat exchanger area for the reboiler (m2)
3
energy cost ($10 /year) B ) bottoms flow rate in the column (mol/s)
conventional 502.9 425.4 361.9 292.8 256.3 195.7 D ) distillate flow rate in the column (mol/s)
reactive column 150.2 136.8 122.8 105.0 92.1 70.0
TAC ($103/year) DC ) diameter of the column (m)
conventional 933.4 797.7 689.9 580.0 514.2 399.9 DR ) diameter of the reactor (m)
reactive column 305.9 271.4 242.3 210.5 190.9 152.7 EF ) activation energy of the forward reaction (cal/mol)
ER ) activation energy of the reverse reaction (cal/mol)
7. Conclusion F ) effluent flow rate from the reactor (mol/s)
Fi ) feed flow rate on tray i (mol/s)
The economically optimum steady-state designs of a
F0j ) fresh feed flow rate of reactant j (mol/s)
multiunit reactor/column/recycle process and a reactive
kF ) specific reaction rate of the forward reaction
distillation column are compared for different values of (kmol‚s-1‚kmol-1)
chemical equilibrium constants. Identical production kR ) specific reaction rate of the reverse reaction
targets and prices are used. Three optimization vari- (kmol‚s-1‚kmol-1)
ables are used to find the optimal results for the two KEQ ) equilibrium constant
systems. The conventional system has the reactor LC ) length of the column (m)
holdup, the reactor temperature, and the composition Li ) liquid flow rate from tray i (mol/s)
of reactant B in the reactor as optimization variables, LR ) length of the reactor (m)
whereas the reactive column configuration has the LR ) liquid flow rate in the rectifying section (mol/s)
number of separation stages, the number of reactive MB ) liquid holdup in the column base (mol)
stages, and the column pressure as optimization vari- MD ) liquid holdup in the reflux drum (mol)
ables. Mi ) liquid holdup on tray i (mol)
As (KEQ)366 becomes larger, the capital cost and the Mw ) molecular weight of all species in the mixture
energy cost decrease for both configurations. These (g/mol)
decreases result in a decrease of the TAC for both NF ) feed tray in the column
systems. When compared for different kinetic cases, NR ) number of rectifying trays
reactive distillation columns have lower TACs than NRX ) number of reactive trays
reactor/column/recycle systems for all values of (KEQ)366. NS ) number of stripping trays
Thus, eliminating a reactor and a column with heat NT ) number of trays in the column
exchangers reduces costs by a factor of 2-3. NTmin ) minimum number of trays in the column
Ideal phase equilibrium is assumed in this study so P ) column pressure (bar)
that a clean comparison can be made and the system- S
Pi,j ) vapor pressure of component j on tray i (bar)
specific complexities of azeotropes do not cloud the R ) ideal-gas-law constant (cal‚mol-1‚K-1)
fundamental issues. The results of this study should Rj ) rate of reaction for component j (mol/s)
provide some general insights and guidance related to RR ) reflux ratio in the column
the differences between the two alternative flowsheets RRmin ) minimum reflux ratio in the column
that arise because of fundamental structural differences Ti ) column temperature on tray i (K)
and limitations. TR ) temperature of the reactor (K)
In this study, we assume that relative volatilities are UC ) overall heat-transfer coefficient in the condenser
constant and fairly large (R ) 2). Studies are underway (kJ‚s-1‚K-1‚m-2)
to explore the effects of changes in the relative volatili- UR ) overall heat-transfer coefficient in the reboiler
ties on the two flowsheets. Two types of changes will (kJ‚s-1‚K-1‚m-2)
be considered: (1) The relative volatilities between Vi ) vapor flow rate from tray i (mol/s)
adjacent products and reactants (RCA and RBD) will be VNT ) molar flow rate from the top of the column (mol/s)
reduced. We expect that this will increase the energy VR ) molar holdup of the reactor (mol)
requirements in both flowsheets and increase the num- VS ) vapor boilup (mol/s)
ber of trays in the distillation columns of the conven- xB,j ) bottoms liquid composition of component j
tional flowsheet and the number of stripping and xD,j ) distillate liquid composition of component j
rectifying trays in the reactive distillation column. There xi,j ) liquid composition of component j on tray i
will probably be little impact on the optimum reactor yB,j ) bottoms vapor composition of component j
temperature in the conventional process. There should yi,j ) vapor composition of component j on tray i
be little effect on the optimum pressure of the reactive zj ) mole fraction of component j in the reactor
distillation column. However, the number of reactive zi,j ) mole fraction of component j in feed tray i
stages might change. (2) The relative volatilities will Greek Symbols
be made temperature-dependent (decreasing with in- Rj ) relative volatility of component j
creasing temperature). This should emphasize the im- βpay ) payback period (year)
portant difference between the two processes. Specifi- ∆HV ) average heat of vaporization (cal/mol)
cally, in the conventional flowsheet, the reactor and ∆TC ) temperature difference in the condenser (K)
column can operate at different pressures. This is not ∆TR ) temperature difference in the reboiler (K)
true in reactive distillation. λ ) average heat of reaction (cal/mol)
Future work will also compare the dynamics and νj ) stoichiometric coefficient of component j
controllabilities of the two systems. χ ) conversion
Ind. Eng. Chem. Res., Vol. 43, No. 10, 2004 2507

Literature Cited (10) Stichlmair, J.; Frey, T. Mixed-integer nonlinear program-


ming optimization of reactive distillation processes. Ind. Eng.
(1) Stankiewicz, A. I.; Moulijn, J. A. Process intensification: Chem. Res. 2001, 40, 5978-5982.
transforming chemical engineering. Chem. Eng. Prog. 2000, 96 (11) Siirola, J. J. An industrial perspective on process synthesis.
(1), 22-34. In Foundations of Computer-Aided Process Design; Biegler, L. T.,
(2) Taylor, R.; Krishna, R. Modeling reactive distillation. Chem. Doherty, M. F., Eds.; AIChE Symposium Series 304; American
Eng. Sci. 2000, 55, 5183-5229. Institute of Chemical Engineers (AIChE): New York: 1995; pp
(3) Agreda, V. H.; Partin, L. R.; Heise, W. H. High-purity 222-233.
methyl acetate production via reactive distillation. Chem. Eng. (12) Chiang, S.; Kuo, C.; Yu, C.; Wong, D. S. H. Design
Prog. 1990, 86, 40-46. alternatives for the amyl acetate process: coupled reactor/column
(4) Doherty, M. F.; Malone, M. F. Conceptual Design of Distil- and reactive distillation. Ind. Eng. Chem. Res. 2002, 41, 3233-
lation Systems; McGraw-Hill: New York, 2001. 3246.
(5) Sharma, M. M.; Mahajani, S. M. Industrial applications of (13) Stitt, E. H. Reactive distillation for toluene disproportion-
reactive distillation. In Reactive DistillationsStatus and Future ation: A technical and economical evaluation. Chem. Eng. Sci.
Directions; Sundmacher, K., Kienle, A., Eds.; Wiley-VCH: New 2002, 57, 1537-1543.
York, 2003; pp 3-29. (14) Luyben, W. L. Economic and dynamic impact of the use of
(6) Ciric, A. R.; Gu, D. Synthesis of nonequilibrium reactive excess reactant in reactive distillation systems. Ind. Eng. Chem.
distillation processes by MINLP optimization. AIChE J. 1994, 40, Res. 2000, 39, 2935-2946.
1479-1487. (15) Douglas, J. M. Conceptual Process Design; McGraw-Hill:
(7) Cardoso, M. F.; Salcedo, R. L.; Feyo de Azevedo, S.; Barbosa, New York, 1988.
D. Optimization of reactive distillation processes with simulated (16) Sneesby, M. G.; Tadé, M. O.; Datta, R.; Smith, T. N. ETBE
annealing. Chem. Eng. Sci. 2000, 55, 5059-5078. synthesis via reactive distillation. 1. Steady-state simulation and
(8) Jackson, J. R.; Grossmann, I. E. A disjunctive programming design aspects. Ind. Eng. Chem. Res. 1997, 36, 1855-1869.
approach for the optimal design of reactive distillation columns. Received for review November 13, 2003
Comput. Chem. Eng. 2001, 25, 1661-1673. Revised manuscript received February 25, 2004
(9) Seferlis, P.; Grievink, J. Optimal design and sensitivity Accepted March 5, 2004
analysis of reactive distillation units using collocation models. Ind.
Eng. Chem. Res. 2001, 40, 1673-1685. IE030832G

You might also like