Vous êtes sur la page 1sur 12

Journal of Process Control 24 (2014) 113124

Contents lists available at ScienceDirect

Journal of Process Control


journal homepage: www.elsevier.com/locate/jprocont

Dynamics and control of benzene hydrogenation via reactive


distillation
Vishal Mahindrakar a , Juergen Hahn a,b,
a
b

Department of Chemical and Biological Engineering, Rensselaer Polytechnic Institute, Troy, NY 12180, United States
Department of Biomedical Engineering, Rensselaer Polytechnic Institute, Troy, NY 12180, United States

a r t i c l e

i n f o

Article history:
Received 14 October 2013
Received in revised form 3 January 2014
Accepted 7 January 2014
Available online 28 February 2014
Keywords:
Reactive distillation
Packed column
Dynamic modeling
Feedback control
Feedforward control

a b s t r a c t
This work develops a dynamic, rst principles-based model of a reactive distillation column used for benzene hydrogenation of a reformate stream and investigates different control structures for this process.
The model is used initially to develop and evaluate a feedback control strategy which provides good regulatory performance for small disturbances, however, it tends to be sluggish for signicant disturbances
in the feed composition. In order to address this point, adding a feedforward controller to the feedback
structure has also been investigated. However, the feedforward controller can only be implemented if
composition measurements of the feed are taken. As online composition measurements are expensive
in practice, several different scenarios have been investigated where samples of the feed are taken and
subsequently analyzed in a lab, as represented by measurement time delays. Simulation results show
that adding feedforward control to the feedback scheme can be very benecial for this process, however,
this is only the case if the composition disturbance measurements do not involve a signicant time delay.
2014 Elsevier Ltd. All rights reserved.

1. Introduction
Automotive emissions are a signicant contributor to poor air
quality [1]. As such, specications for automobile fuels obtained
from petroleum have received increasing levels of attention from
the Environmental Protection Agency (EPA). Benzene is one of the
compounds that is regulated as it is a carcinogen and the EPA
requires all reners to limit the amount of benzene in gasoline to
0.62 vol% [2]. While benzene in the gasoline pool results from a variety of sources, the main contributor is the reformer unit resulting
in signicant amounts of benzene present in reformate streams.
As the reformate stream is used to boost octane rating, there are
economic objectives that have to be taken into account while complying with environmental regulations.
One option to remove benzene is to hydrogenate in the presence
of a catalyst. However, a problem arises as the catalyst used for
the reaction is not exclusively selective for benzene, and toluene,
which is present in the reformate stream in considerable quantities,
will also be hydrogenated. Toluene hydrogenation is undesirable as

Corresponding author at: Center for Biotechnology and Interdisciplinary Studies,


Rensselaer Polytechnic Institute, Troy, NY 12180, United States.
Tel.: +1 518 276 2138; fax: +1 518 276 3035.
E-mail address: hahnj@rpi.edu (J. Hahn).
0959-1524/$ see front matter 2014 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.jprocont.2014.01.005

toluene has a high octane rating (RON) and should be retained in


the nal product.
Benzene (100 RON) + 2H2 cyclohexane (83 RON)

(1)

Toluene (120 RON) + 3H2 methylcyclohexane (75 RON)

(2)

In order to avoid problems related to the selectivity of the catalyst, the reformate stream is split into light and heavy components
in the conventional process (Fig. 1a). As benzene is a reasonably
light component of this mixture, it is mostly concentrated in the
distillate, and accordingly, is hydrogenated before being sent to
the gasoline pool. The downside of this process is that a high capital investment is needed. Reactive distillation (Fig. 1b) offers an
alternative route for solving this problem. By combining reaction
with separation it is possible to selectively react one component in
a specied region of the column while suppressing unwanted reactions of other components. Furthermore, additional savings can be
achieved as the heat of reaction can directly be used for separation
of the mixture.
While reactive distillation (RD) can have signicant advantages
over traditional designs, there are also challenges that need
to be considered. The simultaneous presence of reaction and
separation phenomena can result in complex dynamic behavior.
Combining reaction and separation into a single vessel results in
fewer manipulated variables, thus increasing interactions between
control loops [3]. RD columns have been observed to be very
sensitive to changes in feed concentration. This is a crucial aspect

114

V. Mahindrakar, J. Hahn / Journal of Process Control 24 (2014) 113124

Notation
a
A
CA
dp
ds
D
Ea
F
hi
Hlj
Hvj
HETP
k
K
KA
KH
L
M
Mlj
Mvj
N
P
P0,j
Pj
Q
R
Rgas
Rev
rxn
s
T0
Tj
V
u
x
y
y*
z

geometric surface area of packing per unit volume


(m2 m3 )
cross sectional area of column
concentration (mol m3 )
packing particle diameter (m)
column diameter (m)
distillate ow rate (mol s1 )
reaction activation energy (J mol1 )
feed ow rate (mol s1 )
total liquid holdup based on empty column
(m3 m3 )
molar enthalpy of liquid stream on stage j (J mol1 )
molar enthalpy of vapor stream on stage j (J mol1 )
height equivalent to a theoretical stage (m)
reaction rate constant (mol s1 kg1 )
wall factor
reaction adsorption coefcient (m3 mol1)
reaction adsorption coefcient (m3 mol1 )
liquid ow rate (mol s1 )
mass holdup (kg)
liquid molar holdup on stage j (mol)
vapor molar holdup on stage j (mol)
number of stages
pressure (Pa)
dry column pressure drop across stage j (Pa)
irrigated column pressure drop across stage j (Pa)
external heat energy input (J)
reux ratio
gas constant (J mol1 K1 )
vapor Reynolds number
reaction rate (mol s1 kg1 )
Laplace variable
reaction reference temperature (K)
temperature on stage j (K)
vapor ow rate (mol s1 )
specic liquid load (m s1 )
liquid mole fraction
vapor mole fraction
equilibrium vapor mole fraction
feed mole fraction

Greek letters

packing void fraction


lj,i
liquid fugacity coefcient of component i on stage j
vapor fugacity coefcient of component i on stage j
vj,i
cat




c

catalyst density (kg m3 )


Murphree efciency
relative gain
relative gain array
transfer function time constant (s)
controller design parameter (s)
transfer function time delay (s)
resistance coefcient

Subscripts
i
component index
j
stage index

for benzene hydrogenation as the concentration of some of the


main components in the feed can vary by 50% or more due to
disturbances upstream from the column [4]. The importance of
addressing these disturbances is increased by the fact that changes

in the feed happen on a daily basis and a column operating under a


feedback control can take several hours to return to an acceptable
steady states. This paper investigates these points by developing
a detailed dynamic model, studying the dynamic behavior in
simulations, and developing a control scheme. Furthermore, the
possibility of implementing a feedforward control scheme, in
addition to a feedback one, is investigated where it is taken into
account that feed composition measurements may involve time
delays if the measurements are taken as samples analyzed in a lab.
The outline of this paper is as follows. A literature review is
presented in the following subsection and Section 2 presents preliminary information. A detailed description of the model and
control structure is presented in Section 3. Section 4 discusses column responses to a series of commonly occurring disturbances.
Conclusions are given in Section 5.
1.1. Literature review
Reactive distillation has received a lot of attention as part of
process intensication efforts in the last couple of decades. Employing reactive distillation can result in energy savings as the heat of
reaction is directly used for separation of the mixture. Harmsen
[3] has reviewed commercial applications of reactive distillation.
Reactive distillation systems have been shown to reduce variable
cost, capital expenditure and energy requirements by 20% or more
for some processes [3]. Also, since the heat of reaction is used for
evaporation in a column, increased reaction rates can results in
increased evaporation rates without signicant changes of the temperature. Thus, reactive distillation columns have been found to
be less susceptible to runway behavior than conventional reactors
[3]. Reactive distillation models have been surveyed extensively by
Taylor and Krishna [5] and several articles describing dynamic models, and control structures [69] are available. A variety of different
applications of reactive distillation in reneries have been reported,
such as processes involving ethers (MTBE, ETBE, and TAME [10]).
Sneesby et al. [1113] have developed dynamic models for ETBE
and MTBE, and also made general recommendations for control
system design. Different control strategies for MTBE reactive distillation columns were highlighted by Bartlett and Wahnschafft
[14]. A number of authors have also explored the dynamics and
control for reactive distillation of TAME [1518]. However, despite
these extensive efforts on reactive distillation in general, no papers
on benzene hydrogenation via reactive distillation can be found
in the open literature. This situation is especially peculiar as benzene hydrogenation is an important step in a renery and several
RD columns used for benzene hydrogenation are in operation in
reneries throughout the world.
2. Preliminaries
This section reviews preliminary information needed for the
remainder of the paper. Section 2.1 reviews existing modeling
approaches for reactive distillation columns, some of which will
be used in this work. Existing control strategies for reactive distillation columns are discussed in Section 2.2.1 and Section 2.2.3
reviews the principles of feedforward control which will also be
used.
2.1. Packed column modeling
Reactive distillation can be viewed as an extension of conventional packed columns, where some of the packing includes a
catalyst to facilitate a reaction taking place. A number of papers
have discussed modeling of conventional packed columns. The
key methods used for packed columns are equilibrium (EQ) stage
modeling and non-equilibrium stage modeling (NEQ). In EQ stage

V. Mahindrakar, J. Hahn / Journal of Process Control 24 (2014) 113124

115

Fig. 1. Schematics of (a) conventional reaction/separation process and (b) reactive distillation.

models, the vapor and liquid phase are assumed to be in equilibrium. NEQ stage models use rate-based equations to describe
the mass transfer occurring in conventional distillation columns.
NEQ stage models are generally based on the use of rigorous
MaxwellStefan equations for estimating heat and mass transfer
rates across the interface. Many papers have presented EQ models
[6,8,12,13,1921] and NEQ models [16,17,2022] for reactive distillation. Most of the NEQ models developed for reactive distillation
are generally steady state models [2022]. However, Peng et al. [23]
have compared the results of dynamic NEQ and EQ model, and concluded that the results are similar for their case. Contrary to this,
Baur [17] pointed out that the responses from the models may differ
quantitatively and the dynamics are inuenced by column specications. NEQ models are generally more challenging to simulate
[17,23] and require thermodynamic properties for the calculation
of mass transfer coefcients and interfacial areas.
The dynamic behavior of a distillation column is strongly inuenced by uid hydraulics in the column. This is even more so in
reactive distillation columns, as liquid hold-ups, and liquid residence times are important for determining the conversion and
selectivity of the reactive distillation column. Very few dynamic
models consider both liquid and vapor holdup in the columns.
The vapor holdups are generally neglected because of the low
density of vapor in comparison to liquid. Also, considering vapor
holdups leads to additional computational difculties in the model.
As such, most dynamic models consider only dynamic liquid hold
up or in some cases a constant liquid holdup [23]. However, Choe
and Luyben [24] suggest that vapor holdup should be considered
for dynamic models of columns operating at pressures greater
than 510 atm. Equations governing the vapor and liquid ows,
and hold-ups in a packed column have been discussed by many
papers: Bemer and Kalis [25] have given equations for liquid holdup and pressure drops in irrigated columns while Mackowiak [26]
has extensively reviewed methods for determining vapor ow in
packed columns.

multi-loop control system [27]. A major concern with multi-loop


control is the presence of process interactions, i.e., each manipulated variable may affect multiple controlled variables, however,
these interactions are not taken into account for the control
structure design. Multi-loop control systems may not provide
satisfactory control in some scenarios and multivariable control
strategies, such as model predictive control and decoupling control
can provide better control. However, multi-loop control is the
most widely used control strategy for distillation columns because
of its simplicity, both in terms of maintenance and controller
tuning. As no work has been done on modeling and control of
benzene hydrogenation via reactive distillation, this work focuses
on traditional control strategies and advanced control will be
investigated in the future.
The commonly used control structures for reactive distillation
are similar to those of conventional distillation columns. Skogestad
and co-workers [2830] have discussed the selection of controlled
and manipulated variables. Most distillation columns generally use
either a 4 4 control or a 5 5 control structure. These congurations refer to the number of manipulated variables and the number
of controlled variables used. Eq. (3) presents the set of manipulated
variables u and controlled variables y that are used in 5 5 systems
in no particular order of correlation; see Fig. 2 for an illustration
of the variables. Columns having a 4 4 control system frequently

2.2. Control structure


2.2.1. Feedback control
Reactive distillation columns are systems with multiple inputs
and multiple outputs (MIMO). One approach to deal with MIMO
systems is to treat the control problem as separate individual
loops, i.e., assume that each manipulated variable affects only one
particular controlled variable and design a controller for each loop
separately. This type of control structure is also referred to as a

Fig. 2. Conventional distillation column.

116

V. Mahindrakar, J. Hahn / Journal of Process Control 24 (2014) 113124

do not use the pressure at the top of column, P1 , as a controlled


variable [30].

L1

VN , QN

u=
D
L
N
V2 , QD

x
1
x
N

y=
M1

MN

(3)

P1

2.2.2. Relative gain array


An important general problem for a multi-loop control structure
is to pair the controlled variables and the manipulated variables.
Incorrect pairing may result in poor control performance and interactions among controlled variables. One way to determine the
pairing is Bristols relative gain array (RGA) [42]. Bristol developed
the RGA as a systematic approach to measure the process interactions and recommend an effective pairing of manipulated and
controlled variables. The relative gain between a controlled variable
yi and manipulated variable uj is dened as follows [27]:
(yi /uj )u
(yi /uj )y

open-loop gain
closed-loop gain

11


21
=
.
..

The 5 5 system shown in Fig. 2 is generally used for columns


with a total condenser. For these columns, the pressure is typically controlled by manipulating the condenser heat removal as
the condenser temperature and the column pressure are directly
linked for a total condenser. Partial condensers are used when there
are very light components in the column feed that would require
a high column pressure or a low condenser temperature. In a column with a partial condenser, vapor is removed from the condenser
as a vapor stream. The pressure of the column is strongly inuenced by the outlet vapor stream ow rate. Luyben [31] and Hori
and Skogestad [32] have discussed control structures for columns
with partial condensers. Sloley [33] has extensively reviewed control strategies that can be used for columns based on the type of
condenser.
Recycle loops in chemical plants are known to signicantly alter
the control and dynamics of process networks [3436]. Recycle
streams increase the overall time constants, thus slowing down
its overall response [34]. Designing a control structure for process
networks with recycle can be challenging because recycle streams
can induce a time scale separation, where the dynamics of the process evolves at a fast time scale, and the dynamics of the overall
process with recycle at a slower time scale [37]. Dynamical analysis and control of such process networks with recycle have received
considerable attention [37]. One of the rst design procedures was
proposed by Buckley [38] and has been widely used in industry
for many years [39]. The rst step is to design a control structure that handles the inventory of the entire process (liquid levels
and gas pressures). This hydraulic structure provides smooth
ow rate changes. Fast-acting proportional-only level controllers
provide the most simple and most effective way to achieve this
ow smoothing [39]. The second step is to close the product quality loops. These loops typically use slower proportionalintegral
controllers to hold product streams as close to the specication
as possible. Larsson and Skogestad [40] and Luyben et al. [41] have
discussed plant-wide controller design procedures for a large number of measurements and control loops. When applied to a smaller
single unit scale, they mirror the steps that have been mentioned
above.

ij 

The relative gains are arranged to form the matrix

(4)

n1

12

1n

22

2n

..
.

..
.

..
.

n2

nn

(5)

Controlled and manipulated variables are paired such that the


corresponding relative gains are positive and as close to one as
possible. While the above denition of RGA may seem difcult to
estimate directly for real systems, the RGA can be determined from
an open-loop gain matrix. The procedure for computing the RGA
from an open-loop gain matrix, K, is given by
H = (K 1 )

=K H

(6)
(7)

where denotes the Schur product (element by element multiplication) [27].


The controllers used for feedback control loops are generally
controllers of PID-type, where PI controllers are the most commonly used ones. Luyben [31] has pointed out that ow controllers
that regulate the inventory of a column, e.g., the distillate and bottom ow, should be proportional-only controller as the inventory
in the column is sufciently large to overcome the effect of offsets
that may occur due to the use of proportional-only controllers.
2.2.3. Feedforward control
Feedback control does not take corrective action until after
deviations in the controlled variables occur. As the effects of feed
disturbances will only be detected after a while, this lack of predictive control can limit the overall column performance, especially if
the column includes large time constants or time delays. One option
is to also include feedforward control in addition to feedback control in a control structure. Feedforward control systems measure
the disturbance variables and take corrective action before upsets
of the controlled variables can be recorded. The main disadvantage of feedforward control is that the disturbance variable must
be measured online which is not always feasible, physically or for
economic reasons.
The basic idea for feedforward controller is to measure the
disturbance affecting the system and compute a change of the
manipulated variable such that the effect of the disturbance on
the controlled variable is canceled by the change of the manipulated variable [44]. Feedforward controller designs are thus based
on process models. A feedforward controller transfer function Gf
for a system can be given by [27]
Gf =

Gd
Gp Gmf

(8)

where Gd is the disturbance transfer function, Gp is the process


transfer function, and Gmf is the disturbance sensor/transmitter
transfer function. Sometimes modications to the control law
shown in Eq. (8) need to be made to ensure that the resulting
controller is realizable [27].
Often, the dynamics between a process and a disturbance
are neglected and a simple static feedforward controller may be
designed if the responses are satisfactory. A static feedforward controller is given by the ratio of gains of disturbance, process, and
measurement transfer functions [44]
Gf =

Kd
Kp Kmf

(9)

where Kd , Kp , and Kmf are gains of the disturbance, process, and


measurement transfer functions.

V. Mahindrakar, J. Hahn / Journal of Process Control 24 (2014) 113124


Table 1
Feed composition for the RD column.
Component
n-Butane
n-Pentane
2,3-Dimethylpentane
3-Methylpentane
n-Hexane
Benzene
Cyclohexane
3-Methylhexane
2,4-Dimethylpentane
n-Heptane
Toluene
m-Xylene
Cumene
Hydrogen
Methylcyclohexane

Mole fraction
C4 H10
C5 H12
C7 H16
C6 H14
C6 H14
C6 H6
C6 H12
C7 H16
C7 H16
C7 H16
C7 H8
C8 H10
C9 H12
H2
C7 H14

0.0126
0.0961
0.0116
0.0587
0.0350
0.0826
0.0000
0.0233
0.0234
0.0098
0.2814
0.2063
0.1594
0.0000
0.0000

Feedforward control depends on the accuracy of the disturbance


measurement and on the accuracy of the model describing the process. As some inaccuracies cannot be avoided, feedforward control
by itself would often result in an offset. As a consequence, feedforward control is commonly combined with feedback control in
practice [43].
3. RD column design and control
3.1. Development of dynamic model
The RD column used in this benzene hydrogenation study is
a packed column with a throughput of 200,000 lb/h. The reformate stream enters the column as feed which is processed into
benzene-free lights and heavy stream. The feed stream has 15 components that need to be modeled and the feed composition is given
in Table 1. The column has 70 theoretical stages which includes
a partial condenser and a reboiler. The column stages have been
numbered from top to bottom in this investigation. The rst stage
is the reux drum and the last stage is the reboiler drum. The column model has 10 catalyst stages at the top, i.e., stage 2stage 11.
The feed to the column is added at stage 30. The reformate stream
has a nominal benzene concentration of 6.0 vol% which is common
in reneries [4]. It is expected that the feed benzene concentration may be as high as 11.0 vol% [4]. Hydrogen to the column is
fed at stage 29 along with the unreacted hydrogen that is recycled
from the partial condenser. The column is required to meet EPA
specication of 0.62 vol% maximum benzene concentration at the
outlet during regular operation. Note that the outlet benzene concentration throughout this investigation refers to the total volume
percentage (vol%) taken over all the liquid streams (both distillate
and bottom) exiting the unit.
The RD column is a packed column where a section of the column is lled with catalyst. Industrial data regarding the packing
details and type of catalyst used are unavailable as these are usually
kept a trade secret. Also, no publications on reactive distillation for
benzene hydrogenation are available in the open literature. Thus,
in the absence of any sort of information regarding the packing,
a standard packing size (25 mm pall rings) and catalyst size has
been used in this work. The packing has been treated as equivalent to theoretical trays and a height equivalent to theoretical
packing (HETP) of 0.45 m has been used throughout the column.
The diameter of the column has been determined by assuming
that the vapor velocity reaches a maximum of 80% of the ooding
limit [26]. Based on the vapor ow estimated in the column, the
diameter of the column was estimated to be 2.8 m. Two different
modeling methods are commonly used for each stage of a column:

117

equilibrium-based (EQ) stages and non-equilibrium-based (NEQ)


stages. NEQ models include more detail, but some of the model
parameters are usually not well known and it is unclear if NEQ
models provide a more accurate description than EQ-based models. As such an equilibrium-based modeling approached is used
in this work. Reaction kinetics for the catalyst section have been
taken from Toppinen et al.s [45] work on hydrogenation of benzene and other alkyl benzenes. Benzene hydrogenation columns
operate at a relatively high pressure of 8 atm, and hence variable
vapor holdups have been taken into account. Also, the feed stream
has 15 components that need to be modeled, and some of these
components are bound to have low concentrations in some stages
of column. If a dynamic model considers no vapor hydraulics, then
the vapor ow rates are also dependent on the molar balances. This
creates a problem for the initialization of the model, as the stark differences in the concentration of individual components may lead
to inaccurate estimation of ow rates. In the presence of vapor
hydraulics, the vapor ow rates are governed by the hydraulics
facilitating initialization of the model. Equations governing liq
uid and vapor hydraulics have been adopted from Mackowiaks
compilation [26] on packed bed uid dynamics. The reux drum
and the reboiler drum have been sized to have a residence time
for the liquid of approximately 5 min when the vessels are 50%
full, based on the total amount of liquid entering or leaving the
vessels. Not all of the hydrogen fed to the column will react due
to disturbances in the feed composition. However, hydrogen is
an expensive resource and thus, almost all the unreacted hydrogen is recycled as vapor outlet stream of the partial condenser.
A recycle ratio of 0.99 has been used for the column. The equations of the model as well as the nomenclature can be found in
Appendix A.

3.2. Control structure


3.2.1. Selection and pairing of controlled and manipulated
variables
The set of controlled variables and manipulated variables need
to be identied, in order to design a control structure. A degree
of freedom analysis gives the following set of seven manipulated
variables that can be used for control:

V1

L
N

u = FH2

Q1

QN

(10)

When compared with (3), Eq. (10) has two additional manipulated variables: the reux drum vapor stream ow rate V1 , and
the fresh hydrogen feed ow rate FH2 . This is due to the presence of very light gases such as hydrogen in the system which
has a very low bubble point temperature. Condensing hydrogen
is economically infeasible, which necessitates the need for a partial condenser with a liquid outlet stream and a vapor outlet
stream.
The rst four manipulated variables of u are ow rates and
hence, they are part of the control loops that regulate the inventory of the column. The inuence of these variables on the

118

V. Mahindrakar, J. Hahn / Journal of Process Control 24 (2014) 113124

corresponding controlled variables is straightforward and the


pairing can be performed as follows:

(11)
Even though the objective of an RD column is to maintain the
purity and conversion of the product streams, RD control is based
on temperature points instead of composition. This is because
composition analyzers are expensive to purchase and have high
maintenance costs [47]. They also introduce a delay in measurements if chromatographic methods are used. Temperature sensors
are inexpensive, reliable and introduce small measurement lags.
Based on a degrees of freedom analysis, three temperature points
need to be selected for the remaining three manipulated variables. One of these control points should be somewhere above the
feed and one should be somewhere below the feed. One temperature control point is selected at the top of the column to have
a measurement related to the top product that is located above
the reactive zone. Another of the temperature measurements is
selected approximately halfway between the feed and the bottom
of the column. Since the feed is at stage 30, a temperature measurement at stage 55 is a reasonable choice and has been found
to be sensitive to changes in the manipulated variables. A third
temperature control point needs to be xed at some point within
the column. Hori and Skogestad [46] and Luyben [47] have listed a
number of criteria for selecting the tray at which a temperature sensor should be placed for column control. Conventional techniques
are based, among others, on the slope of the temperature prole,
sensitivity to changes in manipulated variables, SVD analysis, temperature invariance with changes in feed composition. However,
the temperature prole in the reactive zone of the column may
affect the outcomes of these techniques. Therefore a more general approach has been adopted here for selecting the temperature
control tray. The following three criteria have been used:
(i) Avoid trays near the feed tray: the temperature prole near
the feed tray is generally inuenced by the enthalpy of the
feed to the column and may not be as sensitive to changes in
the manipulated variable.
(ii) Avoid trays near the top or the bottom of the column: since
the distillate temperature and one temperature measurement
below the feed have already been used as controlled variables,
any temperature measurement near the top or bottom will be
highly correlated with already selected measurements.
(iii) Avoid the catalyst zone: reactions occurring in the catalyst
zone are exothermic and this affects the temperature of the catalyst stages. A temperature control point should not be selected
in this zone because the temperature is affected by reaction
kinetics in addition to the regular dynamics due to separation.
Based on these criteria the third temperature control point was
chosen to be between the catalyst zone and the feed state. Stage 19
was considered to be a good temperature control point and showed
signicant sensitivity to step changes in the manipulated variables.
Thus, T19 was chosen as the third controlled variable.
The pairing of the remaining manipulated variables was done
via RGA analysis. Step input changes were given to the manipulated
variables of the model with some of the control loops open. Control
loops corresponding to the controlled variables liquid distillate (D),
reux drum vapor ow (V1 ), bottom ow (LN ), and fresh hydrogen
feed (FH2 ) are closed for determining the gain matrix. Simulated data obtained from this model involving partially open-loop

Fig. 3. Schematic of feedback and feedforward control structure for the RD column.

control was tted to transfer functions which were estimated using


the MATLAB system identication toolbox. Most of the responses
were tted to rst order plus time delay (FOPTD) transfer functions. Some of the responses were tted to second order transfer
functions in order to obtain a better t and some of these responses
also included a lead term. Table 2 shows the computed transfer
functions in response to step changes in the manipulated variables.
Based on these transfer functions, the RGA was determined
for the nominal operating conditions (feed benzene concentration = 6 vol%). The RGA was also computed for the RD column at the
extreme operating conditions, i.e., when the feed benzene concentrations are 3 vol% and 11 vol%, since it is possible that the pairing
may change at different operating conditions. These results have
been presented in Table 3 and it can be seen that the pairing of the
controlled and manipulated variables is unaffected by the investigated changes in operating conditions.
Based on RGA computed at the three operating conditions and
the discussion above, the pairing of manipulated and controlled
variables results in the following:

(12)
Fig. 3 gives a schematic of the control structure used for the
column.

V. Mahindrakar, J. Hahn / Journal of Process Control 24 (2014) 113124

119

Table 2
Open loop transfer functions.
R

QN

T1

Q1
1.165 104 (1 + 6052.2 s)

36.523(1 + 9632.8 s)
1 + 2(0.901)(3548.8) s + (3548.8 s)

1 + 2(0.815)(3850.5 s) + (3850.5 s)

T19

15.972
1 + 4564.7 s

2.524 105
1 + 5817.4 s

T55

10.502
1 + 3593.3 s

1.4123 105
(1 + 3081.3 s)(1 + 2259.4 s)

1.0684 104
1 + 1887.1 s

1.8704 105
(1 + 4555.3 s)(1 + 1834.3 s)
1.210 105
1 + 2(0.923)(3013.9) s + (3013.9 s)

Table 3
RGA for RD column at different feed benzene concentrations of (a) 6 vol%, (b) 3 vol%, and (c) 11 vol%.
(a) Feed benzene concentration of 6 vol%

T1
T19
T55

(b) Feed benzene concentration of 3 vol%

(c) Feed benzene concentration of 11 vol%

QN

Q1

QN

Q1

QN

0.643
0.676
2.319

0.157
7.322
6.165

1.800
5.646
4.846

0.908
0.214
2.122

0.026
6.596
5.570

1.933
5.382
4.449

0.518
0.234
1.752

0.106
5.307
4.413

3.2.2. Feedback controller design


Both the reux drum and reboiler drum need controllers to regulate the ows and maintain specied liquid levels in the vessels.
The reux drum also holds vapor which needs to be regulated such
that the pressure of the column is maintained. Since the stream
ow rates regulate the inventory of the column, P-only controllers
have been used. These three proportional controllers were tuned
via Ziegler Nichols tuning relations. A PI controller was used for
maintaining the reux drum outlet vapor ow rate V1 in order
to avoid an offset in the column pressure at the top (P1 ). PI controllers were also used for the temperature point control loops. All
the PI controllers were tuned using internal model control (IMC)
tuning relations [41]. The transfer functions shown in Table 2 were
used to compute the controller parameters for temperature control
loops. A transfer function (13) was obtained for the fresh hydrogen
feed (FH2 ) control loop by passing step change inputs to the model
without the recycle stream.
GPwithout recycle =

V1
75.265
=
FH2
1 + 678.4 s

(13)

An IMC controller was designed based on transfer function (13)


which was determined for the distillation column without the
recycle loop in accordance with design procedure described for
plant-wide control [3840]. Table 4 shows the controller tuning
parameters that were derived and used for the control loops operating on the model. The values of c were chosen by adjusting the
desired speed of the closed-loop response. The faster the desired
response, the lower the value of c . As faster response can lead
to larger overshoots, c needs to be chosen to achieve a tradeoff between speed of response and potential for overshoot. Chien
and Fruehauf [48] have given the following general guideline to
determine acceptable values of c for FOPTD systems,
> c >

(14)

where
is the time delay. The values of c chosen for the PI controlled loops are listed in Table 4.

benzene concentrations for a signicant period of time. As such it


was hypothesized that adding feedforward action to this structure
will improve the performance. The off-spec concentration of benzene in the product stream can be reduced if the ow of hydrogen
to the column is regulated according to the feed composition.
Based on the disturbance variables, manipulated variables, and
controlled variables, Gp and Gd from Eq. (8) are dened as follows:
Gp =

V1
FH2

(15)

Gd =

V1
zC6 H6

(16)

These transfer functions were determined using the MATLAB


system identication toolbox on simulated data for open-loop step
responses:
Gp =

99.39
1 + 3269.9 s

(17)

Gd =

7.457 104
1 + 3580.2 s

(18)

In order to keep the process realistic, feedforward control with


different levels of measurement delay
mf :
Gmf = e
mf s

(19)

were used for the measurement transfer function. Also, since the
feedforward controller is represented by a lead-lag element, the
controller transfer function was augmented with a lter with a
time constant of 120 s in order to avoid large sudden changes in
the manipulated variable. The resulting dynamic feedforward controller is given by
Gf = 750.3

3.3. Feedforward controller


Most conventional columns use feedback-only control [30].
As reactive distillation columns can be viewed as an extension of conventional columns, the rst approach was to design a
feedback-only control structure, appropriately tune the controllers,
and observe the performance. However, as will be shown in Section
4, feed composition disturbances affect the column performance
adversely, i.e., the product does not meet the specications for

Q1
1.412
4.073
3.661

1 + 3269.9 s
1
1 + 3580.2 s 1 + 120 s

(20)

If a static feedforward control law would be considered then this


would result in
Gf = 750.3

(21)

Fig. 3 shows the control structure used for the column along
with the feedforward controller for controlling the fresh hydrogen
feed.

120

V. Mahindrakar, J. Hahn / Journal of Process Control 24 (2014) 113124

Table 4
Feedback controller settings.
Manipulated variable

Controlled variable

Q1

T1

QN

T19

TN

FH2

V1

D
LN
V1

h1
hN
P1

Kc
1.766 107
c
2.305 108
c
342.15

c
9.013
c
3340
3830
12.5

i (s)

c (s)

Type of controller and tuning method used

1887.1

110

PI (IMC)

5817.4

582

PI (IMC)

3593.3

610

PI (IMC)

678.4

128.7

PI (IMC)

P (Ziegler Nichols)
P (Ziegler Nichols)
P (Ziegler Nichols)

Table 5
Steady state results for inlet and outlet streams.
Stream

Feed

Fresh hydrogen

Distillate

Bottom

Vent

Phase
Temperature (K)
Pressure (kPa abs)
Total ow rate (mol s1 )

Liquid
430.0
797.0
265.0

Vapor
430.0
797.0
65.9

Liquid
293.5
792.4
85.4

Liquid
495.5
801.0
180.0

Vapor
293.5
792.4
3.0

Component

Mole fraction

n-Butane
n-Pentane
2,3-Dimethylpentane
3-Methylpentane
n-Hexane
Benzene
Cyclohexane
3-Methylhexane
2,4-Dimethylpentane
n-Heptane
Toluene
m-Xylene
Cumene
Hydrogen
Methylcyclohexane

C4 H10
C5 H12
C7 H16
C6 H14
C6 H14
C6 H6
C6 H12
C7 H16
C7 H16
C7 H16
C7 H8
C8 H10
C9 H12
H2
C7 H14

Benzene conc. (vol%)

1.26E02
9.61E02
1.16E02
5.87E02
3.50E02
8.26E02

2.33E02
2.34E02
9.77E03
2.81E01
2.06E01
1.59E01

1.00E+00

6.010

4. Investigation and comparison of different control


schemes for the column
4.1. Steady state results
The RD column model is assumed to initially operate at the same
steady state for any of the comparisons of different control schemes
made in this section. This steady state corresponds to nominal feed
(composition and temperature) being fed to the column, where the
benzene concentration is 6 vol%. Table 5 includes the steady state
values of all feed and product streams.
Fig. 4 depicts the prole of benzene and toluene concentration
at steady state in the column. The objective of the RD column is to

Benzene molfraction
Toluene molfraction

0.6

Mole fraction

0.5
0.4
0.3

Catalyst
Zone

0.2
0.1
0

10

20

30
40
Stage number

50

Fig. 4. Proles of benzene and toluene in RD column.

60

70

3.87E02
2.97E01
3.58E02
1.82E01
1.08E01
1.09E02
2.45E01
1.54E02
5.92E02
6.48E05
8.86E07
5.20E14
8.42E18
7.10E03
6.09E06

1.85E28
1.32E17
2.59E11
1.32E09
9.37E09
2.74E05
6.26E05
2.70E02
6.27E03
1.43E02
4.14E01
3.04E01
2.35E01
0.00E+00
1.88E04
0.251 (combined)

1.31E02
2.35E02
1.32E03
5.33E03
2.56E03
2.14E04
4.19E03
1.63E04
8.34E04
5.08E07
4.84E09
8.68E17
8.09E21
9.49E01
5.04E08

react as much benzene as possible while minimizing the amount


of toluene entering the stage containing catalysts packing. It can be
seen that the benzene mole fraction increases along the height of
the column, until the reactive zone, where it decreases due to the
reaction. Very little toluene is present in the catalyst zone of the
RD column and as a result 99.8% of the toluene from the reformate
stream is retained in the bottom stream.
4.2. Feedback controller results
The rigorous model built in gPROMS was augmented with the
control structure shown in Fig. 3. The controllers specied in Table 4
were used for investigation. The model combined with the controllers was subjected to changes in the inputs, representing step
disturbances that occur after 1hr, to evaluate the performance of
the control schemes. The set-points of all the controlled variables
remain unchanged throughout this investigation resulting in a regulatory control problem.
Fig. 5 depicts the benzene concentration in the product for the
column under feedback-only control subjected to step changes in
the temperature of 5 K (Fig. 5(a)) and the feed ow rate of 5%
(Fig. 5(b)). The responses indicate that the effect of the disturbances
on the product benzene concentration is not signicant, i.e., only
small changes can be seen in the benzene concentration and the
concentration stays far below the allowable limit.
One of most common disturbances for the benzene hydrogenation process is a change in the feed composition. The benzene
concentration in the reformate stream can increase up to a value

V. Mahindrakar, J. Hahn / Journal of Process Control 24 (2014) 113124

a)

Feed temperature -5 K
Feed temperature +5K

b)
Benzene vol%

Benzene vol%

0.26
0.255
0.25

121

0.27

Feed flowrate +5%


Feed flowrate -5%

0.26

0.25

0.24

0.245
0

Time (hours)

10

15

Time (hours)

10

15

Fig. 5. Responses to step changes in (a) feed temperature and (b) feed ow rate.

of 11 vol% [4]. In order to evaluate such a scenario, a step change


was given to the feed benzene composition from 6 vol% to 11 vol%
and the concentrations of all other components were reduced proportionally. The graphs in Fig. 6 labeled feedback-only show the
responses of the outlet benzene concentration and of the other
manipulated variables for a step change in the feed concentration.
The steady state benzene concentration meets the specications.
However, it can be seen that the responses have signicant overshoot and also a large settling time under a feedback-only control
scheme. This situation presents a clear opportunity for feedforward
control in order to minimize the effect of the disturbance on the
controlled variable.
4.3. Comparison of feedback and feedforward/feedback control
schemes
The simulations shown in Fig. 6 for feedback-only control
involve signicant overshoot and as a result a considerable amount
of product does not meet the EPA specications of 0.62 vol% [2] for

the benzene concentration. Use of feedforward control can reduce


the overshoot and settling time. All subgures in Fig. 6 include a
comparison of the responses of the column for feedback-only control and feedforwardfeedback control to a step disturbance in the
feed composition from 6 vol% to 11 vol% benzene. It can be clearly
seen that there is signicant improvement in the response time
and the overshoot when feedforward control is added to the existing feedback control scheme. Additionally, the trajectories of all
manipulated variables also remain within reasonable bounds for
feedforwardfeedback regulatory control.
While Fig. 6 shows the positive impact that the addition of feedforward control can have on the process, the simulation assumed
that concentration measurements are available instantaneously.
This is not a very realistic assumption in practice unless an online
analyzer is used. One commonly used alternative is that samples
from streams are taken and then analyzed in a lab, i.e., measurements will only be available at discrete points in time and with
a certain time delay. As such, it is important to know how much
of an effect measurement time delay has on the performance of

Fig. 6. Responses of controlled and manipulated variables for step change in feed composition: (a) benzene concentration of the product, (b) fresh H2 feed FH2 , (c) condenser
duty Q1 , (d) reboiler duty QN , and (e) reux ratio R.

122

V. Mahindrakar, J. Hahn / Journal of Process Control 24 (2014) 113124


Feedback-only
FF-FB with continuous measurements
FF-FB with sampling time of 15min
FF-FB with sampling time of 30min
FF-FB with sampling time of 45min
FF-FB with sampling time of 60min

Benzene vol%

0.8
0.6

0.4

0.2
0

0.5

1.5

2.5

3.5

4.5

Time (hours)

Fig. 7. Response to step change in feed composition for feedback-only control and
feedforwardfeedback control for ve different sampling times (continuous, 15 min,
30 min, 45 min, 60 min).

the feedforwardfeedback (FFFB) control system. This case has


been analyzed next. Fig. 7 shows a comparison of the responses for
feedback-only control and combined feedforwardfeedback control where the disturbance measurement transfer function Gmf uses
several different time delays corresponding to the sampling times.
As before, the step change disturbance occurs after an hour and the
system is initially at steady state.
It can be clearly seen that longer measurement time delays signicantly degrade the advantages that the addition of feedforward
control has on the performance. In order to quantify the performance of the responses shown in Fig. 7, a measure is dened to
quantify the area between the curve and the EPA specication of
0.62 vol% [2]:

Dev =

H(Bz vol% 0.62) (Bz vol% 0.62) dt

(22)

where H is the Heaviside function dened as

H(x) =

0,

x < 0

1,

x0

(23)

Table 6 shows a comparison of the deviation value computed


from Eq. (22) for different sampling times. Not surprisingly, it can
be concluded that feedforward control reduces the upset condition,
as measured by Eq. (22), for all investigated cases. Similarly, it is also
not surprising that the best performance is achieved for the combined feedforwardfeedback control structure that uses an online
composition analyzer. However, the largest benzene concentration
Table 6
Deviations, as measured by Eq. (22), for (a) dynamic feedforward controllers and (b)
static feedforward controllers.
(a)
Control structure

Dev

% reduct.

Feedback-only
FFFB with continuous measurements
FFFB with sampling time of 15 min
FFFB with sampling time of 30 min
FFFB with sampling time of 45 min
FFFB with sampling time of 60 min

1074
423
559
756
874
917

61%
48%
30%
19%
15%

Dev

% reduct.

Max Bz
vol%
0.85
0.66
0.76
0.97
1.02
0.94

(b)
Control structure
Feedback-only
FFFB with continuous measurements
FFFB with sampling time of 15 min
FFFB with sampling time of 30 min
FFFB with sampling time of 45 min
FFFB with sampling time of 60 min

1074
444
543
756
888
939

59%
49%
30%
17%
13%

Max Bz
vol%
0.85
0.67
0.74
1.04
1.13
1.04

that is occurring at some point during the operation among all cases
is not occurring for the feedback-only control scheme but instead
for feedforwardfeedback control with signicant time delays. It
can be clearly seen from Table 6 that the larger the time delay for
the composition measurement, the less of a benet in the overall reduction of the offspec product can be achieved. At the same
time, the largest deviations from the target are occurring for long
measurement time delays. It is beyond the scope of this study to
evaluate these responses for different design specications. However, in order to put the discussion of the performance for different
measurement delays into a more general perspective, it should be
pointed out that the dominant time constant of the systems is equal
to 1.6 h. It can be concluded that the feedforwardfeedback scheme
is superior to feedback-only control for the case of continuous measurements or measurements with a time delay of up to 15 min,
which corresponds to 15% of the dominant time constant. There is
only a marginal benet to the feedforwardfeedback scheme if the
time delay is 30 min, corresponding to 31% of the dominant time
constant, and it is questionable if there are any benets of including feedforward control if the measurement time delay is 45 min
or more, corresponding to 46% of the dominant time constant.
In addition to investigating a dynamic feedforward controller, a
static feedforward controller has also been investigated. Table 6b
shows a comparison of the deviation values computed using the
static feedforward controller given by Eq. (21) instead of the
dynamic feedforward controller from Eq. (20). The results are close
to those obtained by using a dynamic feedforward controller. In
fact, the graphs of the responses overlap with those depicted in
Fig. 7 for dynamic feedforward control and as such no separate gure for the graphs is included. Thus, from an application point of
view, a simple static feedforward controller could be used instead
of a dynamic feedforward controller without signicant loss of performance.
One last point to consider is that this investigation focused on
step disturbances as these are the most common disturbances for
the scenario investigated in this work. As such, it was appropriate
to model the measurements occurring from lab samples as continuous samples with a time delay instead of using discrete samples
with time delays as the two will return identical results for step
disturbances. However, it should be pointed out that if the benzene
concentration disturbances would have had a different nature than
a step, that it would have been required to use discrete sampling
and time delays. This was not necessary for the cases investigated
in this work, though.
It was one of the goals of this investigation to determine the
benet of using a control scheme that combines feedforward and
feedback control over a feedback-only control scheme. The simulation results indicate that a signicant benet only exists if upsets
in the feed composition can be quickly detected.
5. Conclusions
Benzene hydrogenation via reactive distillation is a process
that has found signicant use in the process industries. However,
no models of this process can be found in the open literature. This paper addresses this point by developing a dynamic
equilibrium-based model for a reactive distillation column used
for the hydrogenation of benzene. Simulations were carried out to
determine transfer functions between manipulated and controlled
variables. Control loop pairing was performed using RGA analysis and the feedback controllers were tuned via IMC tuning (PI)
and Ziegler Nichols tuning (P). a model-based feedforward controller was also designed to reduce upset conditions caused by
disturbances.
Simulations indicate that the column performance for feed temperature and feed ow rate disturbances remains acceptable for

V. Mahindrakar, J. Hahn / Journal of Process Control 24 (2014) 113124

a feedback-only control scheme. However, the column has a signicant settling time for disturbances in the feed composition.
Feedforward control can reduce these upset conditions resulting
from feed disturbances. However, it was shown that the use of a
feedforwardfeedback control scheme is only benecial if the time
delay associated with the feed composition measurement is small.
In summary, this paper (1) presented the rst detailed model of
a reactive distillation column for the hydrogenation of benzene, (2)
designed and evaluated a feedback control scheme for the column,
and (3) investigated the benet of using a feedforward/feedback
control structure for different sampling times of the feed composition measurement.

123

yj,i = yj+1,i (1 ) + yj,i




(A11)

Energy balance
d(Mlj Hlj + Mvj Hvj )
dt

= Vj+1 Hj+1 + Lj1 Hlj1 Lj Hlj Vj Hvj Hrxn,j


(A12)

Flow rate and holdups


Liquid
Mj vollj = hj (HETP)A

(A13)

2/3

hj = 0.34a1/3 uj
Acknowledgment

Vapor

The authors gratefully acknowledge partial nancial support


from the American Chemical Society - Petroleum Research Fund
(Grant PRF# 50978-ND9).

Pj A(HETP)( hj ) = Mvj Rgas Tj


P0,j
HETP

Appendix A.
A.1. Model

Mass balance
(A1)

Component balances
d(Ml1 x1,i + Mv1 y1,i )
dt
n


x1,i = 1;

i=1

= V2 y2,i L1 x1,i V1 y1,i

i : 1 to n 1 (A2)
rxnj,Toluene =

y1,i = 1

(A3)

(A4)

Energy balance

d(Ml1 Hl1 + Mv1 Hv1 )


dt

(A5)

= V2 H0 (L1 + D)Hl1 V1 Hv1

(A6)

d(MlN )

kj,2 KA2 KH2 CAj,2 CHj


(3KA2 CAj,2 + (KH2 CHj )1/2 + 1)
Ea
Rgas

1
1

Tj
T0

(A20)

(A21)



(A22)

= LN1 LN VN

n



n

= Vj+1 + Lj1 Vj Lj + Ahj (HETP)cat

rxnj,i

(A7)

= Fj zj,i + Vj+1 yj+1,i Lj1 xj1,i Vj yj,i Lj xj,i

i : 1 to n 1

(A8)

yj,i = 1

(A9)

i=1

Vaporliquid equilibrium and Murphree efciency

i : 1 to n

i : 1 to n 1

(A24)

(A25)

Energy balance
dt

+ Ahj (HETP)cat rxnj,i

xN,i = 1

d(MlN HlN )

Component balances

dt

= LN1 xN1,i LN xN,i VN yN,i

i=1

i=1

d(Mlj xj,i + Mvj yj,i )

(A23)

Component balances
dt

Mass balance

xj,i lj,i = yj,i


lj,i

(3KA1 CAj,1 + (KH1 CHj )1/2 + 1)

Reboiler

d(MlN xN,i )

Plate j

i=1

dt

Packed section

xj,i = 1;

(A18)

dp a

Mass balance

0 = V2 Hv2 V1 H0 + Q1

5

kj,1 KA1 KH1 CAj,1 CHj

ki,j = ki,0 exp

i : 1 to n

x1,i l1,i = y1,i v1,i

2hj

(A19)

rxnj,Benzene =

Vaporliquid equilibrium

dt

(A17)

Reaction rate

i=1

d(Mlj + Mvj )

(A16)

Pj+1 = Pj + Pj
= V2 L1 D V1

dt

(A15)

(1 ) u2v v
dp K
3

Pj = P0,j 1

Condenser and reux drum

n


150
+ 1.75
Rev

d(Ml1 + Mv1 )

(A14)

(A10)

= LN1 HlN1 (LN + VN )HlN

VN HlN VN HvN + QN = 0

(A26)
(A27)

References
[1] Mobile Source Air Toxics: Control of Hazardous Air Pollutants from Mobile
Sources, United States Environmental Protection Agency, EPA420-R-05-901,
2005, November.
[2] Control of Hazardous Air Pollutants from Mobile Sources: Final Rule to Reduce
Mobile Source Air Toxics, United States Environmental Protection Agency,
EPA420-F-07-017, 2007, February.
[3] G.J. Harmsen, Reactive distillation: the front-runner of industrial process intensication: a full review of commercial applications, research, scale-up, design
and operation, Chemical Engineering and Processing: Process Intensication
46 (9) (2007) 774780.

124

V. Mahindrakar, J. Hahn / Journal of Process Control 24 (2014) 113124

[4] Control of Hazardous Air Pollutants from Mobile Sources: 40 CFR Parts 59, 50,
85 and 86, Environmental Protection Agency, EPA-HQ-OAR-2005-0036, 2006,
February.
[5] R. Taylor, R. Krishna, Modelling reactive distillation, Chemical Engineering Science 55 (22) (2000) 51835229.
[6] K. Alejski, F. Duprat, Dynamic simulation of the multicomponent reactive distillation, Chemical Engineering Science 51 (18) (1996) 42374252.
[7] D.B. Kaymak, W.L. Luyben, Evaluation of a two-temperature control structure
for a two-reactant/two-product type of reactive distillation column, Chemical
Engineering Science 61 (13) (2006) 44324450.
[8] M.C. Georgiadis, M. Schenk, E.N. Pistikopoulos, R. Gani, The interactions of
design control and operability in reactive distillation systems, Computers &
Chemical Engineering 26 (4) (2002) 735746.
[9] P. Kumar, N. Kaistha, Decentralized control of a kinetically controlled ideal reactive distillation column, Chemical Engineering Science 63 (1) (2008) 228243.
[10] G.R. Gildert, K. Rock, T. McGuirk, Advances in process technology through catalytic distillation, in: Proceeding of the International Symposium on Large
Chemical Plants, Antwerpen, 1998, pp. 103113.
[11] M.G. Sneesby, M.O. Tade, T.N. Smith, Two-point control of a reactive distillation
column for composition and conversion, Journal of Process Control 9 (1) (1999)
1931.
[12] M.G. Sneesby, M.O. Tade, T.N. Smith, Steady-state transitions in the reactive
distillation of MTBE, Computers & Chemical Engineering 22 (7) (1998) 879892.
[13] M.G. Sneesby, M.O. Tade, R. Datta, T.N. Smith, ETBE synthesis via reactive distillation. 2. Dynamic simulation and control aspects, Industrial & Engineering
Chemistry Research 36 (5) (1997) 18701881.
[14] D.A. Bartlett, O.M. Wahnschafft, Dynamic simulation and control strategy
evaluation for MTBE reactive distillation, in: Foundations of Computer-Aided
Process Operation, AIChE Symposium Series, vol. 320, Snowbird, 1998, pp.
315321.
[15] N. Sharma, K. Singh, Model predictive control and neural network predictive control of TAME reactive distillation column, Chemical Engineering and
Processing: Process Intensication 59 (2012) 921.
[16] A.M. Katariya, R.S. Kamath, K.M. Moudgalya, S.M. Mahajani, Non-equilibrium
stage modeling and non-linear dynamic effects in the synthesis of TAME
by reactive distillation, Computers & Chemical Engineering 32 (10) (2008)
22432255.
[17] R. Baur, R. Taylor, R. Krishna, Dynamic behaviour of reactive distillation columns
described by a nonequilibrium stage model, Chemical Engineering Science 56
(6) (2001) 20852102.

[18] A. Koodziej, M. Jaroszynski,


W. Saacki, W. Orlikowski, K. Fraczek,
M. Klker,
A. Grak, Catalytic distillation for TAME synthesis with structured catalytic
packings, Chemical Engineering Research and Design 82 (2) (2004) 175184.
[19] R. Jacobs, R. Krishna, Multiple solutions in reactive distillation for methyl tertbutyl ether synthesis, Industrial & Engineering Chemistry Research 32 (8)
(1993) 17061709.
[20] R. Baur, A.P. Higler, R. Taylor, R. Krishna, Comparison of equilibrium stage and
nonequilibrium stage models for reactive distillation, Chemical Engineering
Journal 76 (1) (2000) 3347.
[21] J. Peng, S. Lextrait, T.F. Edgar, R.B. Eldridge, A comparison of steady-state
equilibrium and rate-based models for packed reactive distillation columns,
Industrial & Engineering Chemistry Research 41 (11) (2002) 27352744.
[22] A.P. Higler, R. Taylor, R. Krishna, Nonequilibrium modelling of reactive distillation: multiple steady states in MTBE synthesis, Chemical Engineering Science
54 (10) (1999) 13891395.
[23] J. Peng, T.F. Edgar, R.B. Eldridge, Dynamic rate-based and equilibrium models
for a packed reactive distillation column, Chemical Engineering Science 58 (12)
(2003) 26712680.
[24] Y.S. Choe, W.L. Luyben, Rigorous dynamic models of distillation columns, Industrial & Engineering Chemistry Research 26 (10) (1987) 21582161.

[25] G.G. Bemer, G.A.J. Kalis, A new method to predict hold-up and pressure drop
in packed columns, Transactions of the Institution of Chemical Engineers 56
(1978) 200.
[26] J. Mackowiak, Fluid Dynamics of Packed Columns: Principles of the Fluid
Dynamic Design of Columns for Gas/Liquid and Liquid/Liquid Systems,
Springer, New York, 2010.
[27] D.E. Seborg, D.A. Mellichamp, T.F. Edgar, F.J. Doyle III, Process Dynamics and
Control, Wiley, New York, 2010.
[28] S. Skogestad, M. Morari, Control conguration selection for distillation
columns, AIChE Journal 33 (10) (1987) 16201635.
[29] S. Skogestad, P. Lundstrm, E.W. Jacobsen, Selecting the best distillation control
conguration, AIChE Journal 36 (5) (1990) 753764.
[30] S. Skogestad, Dynamics and control of distillation columns a critical survey,
Modeling, Identication and Control 18 (3) (1997) 177217.
[31] W.L. Luyben, Alternative control structures for distillation columns with partial condensers, Industrial & Engineering Chemistry Research 43 (20) (2004)
64166429.
[32] E.S. Hori, S. Skogestad, Control structure selection of a deethanizer column with
partial condenser, in: Proceedings of European Congress of Chemical Engineering (ECCE-6), Copenhagen, 2007.
[33] A.W. Sloley, Effectively control column pressure, Chemical Engineering
Progress 97 (1) (2001) 3848.
[34] M.M. Denn, R. Lavie, Dynamics of plants with recycle, Chemical Engineering
Journal 24 (1) (1982) 5559.
[35] N. Kapoor, T.J. McAvoy, T.E. Marlin, Effect of recycle structure on distillation
tower time constants, AIChE Journal 32 (3) (1986) 411418.
[36] J. Morud, S. Skogestad, Effects of recycle on dynamics and control of
chemical processing plants, Computers & Chemical Engineering 18 (1994)
S529S534.
[37] A. Kumar, P. Daoutidis, Nonlinear dynamics and control of process systems with
recycle, Journal of Process Control 12 (4) (2002) 475484.
[38] P.S. Buckley, Techniques of Process Control, Wiley, New York, 1964.
[39] W.L. Luyben, Dynamics and control of recycle systems. 1. Simple open-loop
and closed-loop systems, Industrial & Engineering Chemistry Research 32 (3)
(1993) 466475.
[40] T. Larsson, S. Skogestad, Plantwide control a review and a new design procedure, Modeling, Identication and Control 21 (4) (2000) 209240.
[41] M.L. Luyben, B.D. Tyreus, W.L. Luyben, Plantwide control design procedure,
AIChE Journal 43 (12) (1997) 31613174.
[42] E.H. Bristol, On a new measure of interactions for multivariable process control,
IEEE Transactions on Automatic Control AC-II (1966) 133.
[43] F.G. Shinskey, Process Control Systems: Application, Design and Tuning,
McGraw-Hill, Inc., New York, 1990 (Chapter 7).
[44] B.W. Bequette, Process Control: Modeling, Design, and Simulation, Prentice Hall
Professional, New Jersey, 2003.
[45] S. Toppinen, T.K. Rantakyl, T. Salmi, J. Aittamaa, Kinetics of the liquid-phase
hydrogenation of benzene and some monosubstituted alkylbenzenes over a
nickel catalyst, Industrial & Engineering Chemistry Research 35 (6) (1996)
18241833.
[46] E.S. Hori, S. Skogestad, Selection of controlled variables: maximum gain rule and
combination of measurements, Industrial & Engineering Chemistry Research
47 (23) (2008) 94659471.
[47] W.L. Luyben, Evaluation of criteria for selecting temperature control
trays in distillation columns, Journal of Process Control 16 (2) (2006)
115134.
[48] I.-L. Chien, P.S. Fruehauf, Consider IMC tuning to improve controller performance, Chemical Engineering Progress 86 (1990) 3341.

Vous aimerez peut-être aussi