You are on page 1of 12

Chemical Engineering and Processing 50 (2011) 281–292

Contents lists available at ScienceDirect

Chemical Engineering and Processing:


Process Intensification
journal homepage: www.elsevier.com/locate/cep

A control perspective on process intensification in dividing-wall columns


Anton A. Kiss a,∗ , Costin Sorin Bildea b
a
AkzoNobel Research, Development & Innovation, Process Technology ECG, Velperweg 76, 6824 BM Arnhem, The Netherlands
b
University “Politehnica” of Bucharest, Centre for Technology Transfer in Process, Industries, Polizu 1-7, 011061 Bucharest, Romania

a r t i c l e i n f o a b s t r a c t

Article history: During the last decades, process intensification led to major developments also in separation technology.
Received 17 December 2010 Particularly in distillation, dividing-wall column (DWC) is the next best thing as it allows significant
Received in revised form 25 January 2011 energy savings combined with reduced investment costs. However, in spite of these clear advantages
Accepted 28 January 2011
and the steady increase of DWC applications, the spreading of DWC at industrial scale is still limited to
Available online 1 March 2011
only a few companies. One of the major reasons for this status quo is the insufficient insight with regard
to the operation and control of a DWC – this lack of knowledge making most chemical companies reticent
Keywords:
to large-scale implementations. This study gives an overview of the available control strategies for DWC,
Dividing-wall column
Petlyuk
varying from the classic three-point control structure and PID controllers in a multi-loop framework to
Control strategies model predictive control (MPC) and other advanced control strategies (LQG, LSDP, H∞ and ␮-synthesis).
PID The previous studies prove that the DWC is not difficult to control providing that an appropriate control
MPC structure is selected. The available results show that MIMO controllers perform better than multi-loop
LQG PID controllers. However, among the decentralized multivariable PI structured controllers, LSV and DSV
LSDP are the best control structures being able to handle persistent disturbances in reasonably short times.
H∞ All things considered, this study clearly concludes that the DWC controllability is only perceived as a
␮-Synthesis
problem, but in fact there are no real solid grounds for concern.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction appropriate position – basically integrating the Petlyuk system into


a single column shell [1,8–10]. Fig. 1 illustrates the conventional
As a thermal separation method, distillation remains one of the direct and indirect sequences, Petlyuk configuration and DWC sep-
most important separation technologies in the chemical process aration alternatives. Note that using a DWC requires an overlapping
industry. Nevertheless, in spite of its many well-known benefits of the operating windows (e.g. pressure) of the two stand-alone
and the widespread use, one major drawback is the significant columns in a conventional direct or indirect sequence [11].
energy requirements, since distillation can generate more than 50% In a DWC, the feed is introduced into the prefractionator (feed
of plant operating cost. An innovative solution to overcome this side of the column) while a side stream is removed from the main
drawback is using advanced process integration and process inten- column (side stream section). The side stream contains the inter-
sification techniques such as thermally coupled or heat-integrated mediate boiling component, while the lightest component goes
distillation columns [1–6]. overhead in the distillate product and the heaviest component goes
Typically, ternary mixtures are separated via a direct sequence out in the bottoms product. At the top of the dividing-wall section,
(most volatile component is separated first), indirect sequence the liquid coming down from the rectifying section is split between
(heaviest component is separated first) or distributed sequence the two sides, a part of the total liquid being sent to the prefrac-
(mid-split) consisting of two or more distillation columns. These tionator side and the rest to the side stream side of the column.
separation sequences advanced to the Petlyuk configuration [7] Assuming similar hydrodynamic conditions on both sides of DWC,
consisting of two fully thermally coupled distillation columns. One the vapor flow is split proportionally to the cross-sectional area of
practical implementation of the Petlyuk column is the dividing- each side (Fig. 1).
wall column (DWC) that splits the middle section of a single vessel DWC found great appeal in the chemical industry – with Montz
into two sections by inserting a vertical wall in the vessel at an and BASF as the leading companies [12] – because it can separate
three or more components in a single tower – thereby saving the
cost of building two columns and cutting operating costs using a
∗ Corresponding author. Tel.: +31 26 366 1714; fax: +31 26 366 5871. single condenser and reboiler. In fact, using dividing-wall columns
E-mail addresses: Tony.Kiss@akzonobel.com, tonykiss@gmail.com (A.A. Kiss), can save up to 30% in the capital invested and up to 40% in the energy
s bildea@upb.ro (C.S. Bildea). costs [9,13,14], particularly for close boiling-species. Remarkably,

0255-2701/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.cep.2011.01.011
282 A.A. Kiss, C.S. Bildea / Chemical Engineering and Processing 50 (2011) 281–292

A B A

ABC 1 2 ABC 1 2
BC
AB

C C B
Direct sequence Indirect sequence

Liquid split
A

Dividing wall

ABC ABC
1 2 B
B

Prefractionation
section
Main column

Vapour split

C C
Petlyuk configuration Dividing-wall column

Fig. 1. Separation of a ternary mixture via direct and indirect distillation sequences, Petlyuk configuration and dividing-wall column (DWC).

DWC is not limited to ternary separations alone, but it can be used ity in operation. But what is in fact this controllability property
also in azeotropic separations [15], extractive distillation [16], and and why is it so important? Basically, the controllability of a sys-
reactive distillation [14,17,18]. tem denotes the ability to reject the expected disturbances and
Compared to classic distillation design arrangements, DWC to move the system to new operating points using only certain
offers several important benefits [6,9,11,19]: admissible manipulations [28]. Practically, this means that a DWC
should be able to deliver on-spec products in spite of common
• High thermodynamic efficiency due to reduced remixing effects. transitory regimes arising due to planned changes or unexpected
• Lower energy requirements as compared to conventional sepa- disturbances.
ration sequences. Although a large amount of the existing literature focuses on
• High purity for all three or more product streams reached in only the control of binary distillation columns, there are only a limited
one column. number of studies on the control of DWC. The literature review
• Small footprint and low investment due to the reduced number that follows makes a critical overview of the most important DWC
of equipment units. control studies up to date. It should be noted that various authors
• Reduced maintenance costs as compared to traditional distilla- have selected different ternary chemical systems to be separated
tion sequences. and have explored many control structures (the selection and
pairing of manipulated and controlled variables) with different
control objectives and different control algorithms, from propor-
Note however that the integration of two columns into one shell
tional integral derivative (PID) to model predictive control (MPC).
leads also to changes in the operating mode and the controlla-
Some authors control only two compositions, while others control
bility of such an integrated system [20–26]. Therefore, all these
three or even four compositions. Several papers look at inferential
benefits are possible only under the condition that a good control
temperature control instead of – or in combination with – com-
strategy is available and able to attain the separation objectives
position control. Nevertheless, the main conclusion of this study
[45].
is that the dividing-wall distillation column has good controllabil-
ity properties, providing that an appropriate control structure is
2. Problem statement implemented.
The approach we take here is to present previously reported
The advantages of using the dividing-wall distillation for ternary work on the controllability of DWC, by following the historical
separation are the main drivers for commercial implementation development of different control structures. For each paper ana-
– especially for mixtures where the middle-boiling component is lyzed afterward, the novelty and key details of the approach are
in the largest amount [6]. However, there are also major hurdles presented and – when available – the results of the dynamic simula-
[27], such as the concern that the benefits of DWCs are obtained tions proving the performance of the controllers are also discussed
at the cost of lack of controllability and consequently flexibil- and/or shown. Two relevant case studies are also briefly described.
A.A. Kiss, C.S. Bildea / Chemical Engineering and Processing 50 (2011) 281–292 283

reflux (L) and the reboiler duty (Qr ) – or, equivalently, the vapor
boilup rate (V). These variables can be also combined, for exam-
QC ple, instead of the reflux rate L the reflux ratio R = L/D could be
manipulated.
N1 + N2+1 The additional degree of freedom arises from the flow of liquid
between the two sections of the column: at the top of the divided-
N1 + N2+2 L D wall section, the liquid coming down from the rectifying section can
3 .
. be split, in a controlled manner, between the two sides of the wall
βL .
N1 + N2+N3
using a total liquid trap-out tray and sending part of the total liquid
to the prefractionator side and the rest to the side stream side. Thus
1 this internal liquid split (ˇL ) is available for control purposes. Note
N1 + N2+N3 + 1
2 .
that at the bottom of the dividing-wall section, the vapor flow is
.
.
1 4 .
.
split proportionally to the cross-sectional area of each side and the
. S hydrodynamic conditions (e.g. flow resistance). The cross-sectional
F N1
N1 + N2+N3 + N4 area of each side is fixed by the physical location of the wall, and
N1 + N2+N3 + N4 + 1 this is already set at the design stage hence it cannot be changed
N1+1
.
later on, during operation. Because the location of the wall fixes
N1+2
. how the vapor flow splits between the two sides of the column,
. 2 5 .
. the vapor split (ˇV ) variable is not adjustable during operation for
.
. control purposes.
N1 + N2 N1 + N2+N3 + N4 + N5
Typically the distillate (D) and bottoms flow rates (B) are used to
N1 + N2+N3 + N4 + N5 +1 maintain liquid levels in the reflux drum and column base, respec-
.
tively. Moreover, the condenser duty (Qc ) controls the pressure.
6 .
Therefore, the four degrees of freedom left (L, S, Qr , ˇL ) can be
N1 + N2+N3 + N4 + N5 +N6 -1
used to control four variables. Note that in some control struc-
V QB tures, the roles of the flow rates D and B can be exchanged with
that of the flow rates L and V, respectively. Ideally the purities of all
three product streams should be controlled: the amount of inter-
N1 + N2+N3 + N4 + N5 +N6 B mediate component 2 in distillate and bottom streams (xD2 and
xB2 , respectively) and the amounts of light and heavy components
Fig. 2. Schematics of a dividing-wall column, including notation of streams. in the side stream (xS1 and xS3 ). However, the last set of specifi-
cations is usually replaced by the purity of the side stream (xS2 ),
3. Degrees of freedom analysis which means that one degree of freedom can be used to achieve
some other objective, as for example to minimize the energy
In the following sections, the first subscript in the concentration requirements.
notation denotes the stream (D, S or B), while the second subscript
will refer to the components. The components will be labeled 1
(or A), 2 (or B) and 3 (or C), for the light, intermediate and heavy,
4. Overview of DWC control structures
respectively. Fig. 2 illustrates the schematics of a DWC divided into
six sections, while Table 1 provides the flow relationships in the
All the control structures presented hereafter could use or make
DWC based on the mole balance at the interface between sections.
use of an additional optimization loop that manipulates the liquid
Remarkably, the specification of L, V, S, ˇL , ˇV is sufficient to deter-
split in order to control the heavy component composition in the
mine all the flow rates in a DWC, in all sections.
top of fractionator, and implicitly achieving minimization of the
The dividing-wall distillation system has seven degrees of free-
energy requirements. Ling and Luyben [29] have already shown
dom. Six degrees of freedom are usual for a typical distillation
that implicit optimization of the energy usage is achieved by con-
column with side stream, namely: the condenser duty (Qc ), the
trolling the heavy impurity at the top of the prefractionator.
product rates (distillate D, side stream S, bottoms B), the liquid
Many authors analyzed the effect of minimizing the energy
requirements on the controllability properties of the system – as
Table 1 an optimal design might give the highest energy savings but lack
Flow relationships in a DWC, based on the mole balance at the interface between good controllability [30–32]. Rong and Turunen [33] reported a reli-
sections. Specification of L, V, S, ˇL , ˇV is sufficient to determine all flow rates in the
able synthesis method while Gomez-Castro et al. [34] proposed a
DWC.
robust method for the design of distillation sequences with dividing
Section Liquid flow Vapor flow walls and recommended an analysis of both thermodynamic and
Reboiler V6 = V controllability properties. Accordingly, it was concluded that the
From 6 to 2 V2 = ˇV · V6 distillation arrangements (e.g. DWC) other than conventional Pet-
From 6 to 5 V5 = (1 − ˇV ) · V6 lyuk, presented the best values for condition number and minimum
From 2 to 1 V1 = V2 + (1 − q) · F
singular value. As a result, good dynamic closed-loop performance
From 5 to 4 V4 = V5
From (1 + 4) to 3 V3 = V1 + V4 could be expected for these types of thermally coupled distillation
Condenser D = V3 − L systems [34]. Moreover, it was found that thermally coupled distil-
L3 = L lation systems (TCDS) are not only well controllable but sometimes
From 3 to 1 L1 = ˇL · L3
they exhibit dynamic responses that are easier to manage than in
From 1 to 2 L2 = L1 + q · F
From 3 to 4 L4 = (1 − ˇL ) · L3
the case of conventional distillation sequences [31,34,35].
From 4 to 5 L5 = L4 + S Table 2 gives an overview of the main DWC control structures
From (2 + 5) to 6 L6 = L 2 + L 5 reported so far in the open literature, while the next sections pro-
Reboiler B = L6 − V vide a more detailed analysis of each of them.
284 A.A. Kiss, C.S. Bildea / Chemical Engineering and Processing 50 (2011) 281–292

Table 2
Overview of various DWC control structures.

DWC control structure Ternary system References

Three-point control structure Ethanol/propanol/butanol Wolff and Skogestad [36]


Four-point control structure
Three-point control structure with nested loops
Three-point control structure with nested loops Methanol/2-propanol/butanol Abdul Mutalib and Smith [38], Abdul Mutalib
et al. [50]
Three-point control structure with alternative pairing Hypothetical system with constant relative Serra et al. [20,21,22]
volatilities: 1:2.15:4.65
Performance control of prefractionator sub-system using Hypothetical system with constant-relative Halvorsen and Skogestad [39,40]
the liquid split volatility. Relative volatilities: 1:2:4, boiling
points 100, 50, 0 ◦ C
Control of the temperature in the top of the prefractionator Butanol/pentanol/hexanol Adrian et al. [42]
Model predictive control (MPC)
Linear controllability analysis, nonlinear dynamic Pentane/hexane/heptane Bildea [47]
simulations
Control of the temperature in the bottom of the Ethanol/1-propanol/1-butanol Wang and Wong [25]
prefractionator
Temperature–composition cascade
Null space method Hypothetical system with constant-relative Alstad and Skogestad [43]
Self-optimizing control volatility. Relative volatilities: 1:3:9
Feedforward control to reject frequent measurable Benzene/toluene/xylene Ling and Luyben [29]
disturbances
Feedforward and differential temperature control loops Benzene/toluene/xylene Ling and Luyben [41]
LQG/LQR, higher order controllers based on H∞ norm Benzene/toluene/xylene Van Diggelen et al. [45]
␮-synthesis

4.1. Three-point control structure the PI control with dynamic matrix control (DMC), revealing the
better performance of the PI control and the limitations of the
The simplest control structure that can be imagined is just an DMC [22,37]. The authors also pointed out that the performance
extension of the control of a regular distillation column with a side is dependent on the design of the column, but did not elaborate
stream. This is known also as the three-point control structure, and further on this observation.
it is shown in Fig. 3, left. The distillate purity (xD1 ) is controlled by
manipulating the reflux rate (L), the side stream purity (xS2 ) is con- 4.3. Four-point control structure
trolled by manipulating the side stream flow rate (S) and the bottom
purity (xB3 ) is controlled by manipulating the vapor boilup (V). The In addition to the reflux, side stream and boilup that control
three-point structure LSV was suggested by Wolff and Skogestad the concentration of the main components in the product streams,
[36]. They studied the separation of an ethanol/propanol/butanol the liquid split ratio (ˇL ) can be added to the set of manipulated
mixture in a system with 20-tray prefractionator and 40-tray main variables with the goal of controlling the levels of both impurities in
column. The authors analyzed the controllability of the system the side stream. The resulting four-point control structure is shown
using linear tools, concluding that the system is easy to be con- in Fig. 3, right. Wolff and Skogestad [36] tested this structure on the
trolled. Moreover, they also performed dynamic simulations of separation of the same system previously described. Both linear
the nonlinear model, showing that the column handles well dis- and nonlinear tools predicted difficult control. Besides small gain
turbances and some setpoint changes. However, sometimes small from the manipulated variables towards the controlled variables,
changes in the purity setpoints (for example xS2 , from 0.994 to the steady state feasibility space shows regions where no steady
0.996) could not be handled – changes of 100% in L and V being states exists.
required. The most likely reason for this is that the column does
not have enough stages to achieve the degree of separation, at least 4.4. Three-point control structure with nested loops
not without adjusting the liquid (ˇL ) and vapor (ˇV ) split ratios.
Note that the mole fraction xS2 does not define the composition of As an alternative, Wolff and Skogestad [36] switched the V–xB3
the side stream, as it can correspond to several levels of the concen- and S–xS2 control loops to the pairing V–xS2 and S–xB3 , as illustrated
tration of impurities xS1 and xS3 . Although this does not appear to be by Fig. 4, left. Although the linear controllability tool did not predict
a problem in a practical implementation, the authors tried to find control difficulties, the structure proved unworkable under mild
a solution using the remaining available degree of freedom (liquid disturbances. The authors attributed this to the strong nonlinearity
split ratio, ˇL ) hence adding another point to the control structure. of the pairing from V to xS2 . However, it should be observed that the
loops V–xS2 and S–xB3 are nested: a change of the side stream flow
4.2. Three-point control structure with alternative pairing rate is expected to affect first the side stream purity xS2 and only
later the bottom purity xB3 . Therefore, strong interactions between
Serra et al. [20–23] performed a comprehensive study concern- the loops are expected.
ing the operation and controllability of dividing-wall distillation The same control structure was also considered by Abdul
columns. They studied a model system with constant relative Mutalib et al. [38,50]. However, they presented a simulation study
volatility (˛ = 1:2.15:4.65). Several control structures where inves- of the methanol/2-propanol/butanol separation system. Unlike
tigated, each of structure using a different set of manipulated Wolff and Skogestad [36], it was found that the pairing V–xS2
variables to control the products purities. However, the liquid and S–xB3 worked well in the three-composition control structure.
split ratio (ˇL ) was not considered. Different controllability indices Moreover, they switched the reflux and distillate in the level and
were used to assess the performance of the pairing in a three- distillate-concentration loops, as shown in Fig. 4, right. The com-
point control structure. The results showed that the best structure parison of performance between the two control structures, for
was the one previously introduced by Skogestad and Wolff (Fig. 3, setpoint changes and feed disturbances, showed that the system
left). Moreover, Serra et al. [20–22] compared the performance of was indeed controllable [50].
A.A. Kiss, C.S. Bildea / Chemical Engineering and Processing 50 (2011) 281–292 285

Fig. 3. Three-point (left) and four-point (right) control structures.

4.5. Performance control of prefractionator sub-system using the Any amount of heavy component going out the top of the wall
liquid split will end up also in the liquid flowing down in the main column and
thus strongly affecting the purity of the side stream (S). Similarly,
Halvorsen and Skogestad [39,40] pointed out two key tasks that any portion of the lightest component that goes out the bottom of
should be achieved by the prefractionator when the system is facing the prefractionator section will flow up through the side stream
disturbances or setpoint changes are required: section, mostly in the vapor phase, with a reduced effect on the
composition of the side stream. Since the side stream is collected
1. Keep the heaviest component from going out the top of the pre- as a liquid product, it means that any small amounts of light impu-
fractionator section, and rity in the vapor phase will not significantly affect the side stream
2. Keep the lightest component from going out the bottom of the composition. However, even tiny amounts of heavy impurity in the
prefractionator section. liquid phase will greatly affect the side stream composition.

Fig. 4. Three-point control structure with reversed loops (left) or with switched loops pairing (right).
286 A.A. Kiss, C.S. Bildea / Chemical Engineering and Processing 50 (2011) 281–292

PC PC
QC QC
LC LC

3 L 3 L :
D D
CC RC
CC βL
βL

TT
TT
1 4 CC 1 4
S S
F F

FC
2 5 2 5
MPC

TT
6 6 V
CC FC
V QB QB

LC B LC B

Fig. 5. Control structure using the liquid split to control the heavy impurity on the Fig. 7. Model predictive control of a DWC, using temperature measurements.
top of the prefractionator.

4.6. Control structures based on inferential temperature


measurements
Based on these considerations, the authors added a fourth
control loop where the liquid split ratio (ˇL ) is used to control 4.6.1. Controlling the temperature in the top of the
the level of the heavy impurity in the top of the prefractionator prefractionator
(Fig. 5). This approach was successfully used later also by other Adrian et al. [42] reported interesting experimental results
authors [29,41]. Note that the mixture considered for separation concerning the control of a butanol/pentanol/hexanol system.
was a hypothetical constant-relative volatility system. The relative Compared to previous studies, temperature control was used
volatilities were 1:2:4, and the boiling points were 100, 50, 0 ◦ C instead of composition control. The location of the controlled tem-
[39,40]. peratures included the top of the prefractionator, a point above the

Fig. 6. DWC control structure using temperature measurements for the top of the prefractionator (left) or the bottom of the prefractionator (right).
A.A. Kiss, C.S. Bildea / Chemical Engineering and Processing 50 (2011) 281–292 287

PC QC PC
QC

LC LC

3
L L
3 D CC
D
+
TC CC
βL βL
CC

1 1 4 CC
4 FFC S
S
F F

FFC 2 5
2 5
TC

TC
6
6 CC + CC
V
V QB
QB
LC B
CC
LC B
Fig. 9. Four-point control structure using combined feedback–feedforward to
Fig. 8. DWC control structure using a concentration–temperature cascade (right). account for feed flow disturbances.

side draw and the lower part of the column (Fig. 6, left). It should be
also remarked that the heat input (reboiler duty, Qr ) was fixed. An Settling time for +10% persistent disturbance in feed flowrate
alternative configuration was proposed later by Wang and Wong
μ-controller 642
[25], in which one of the controlled temperatures is located in the
bottom of the prefractionator side (Fig. 6, right). LSDP 645
Model predictive control (MPC) was employed by a second con- LQR-IA 510
trol structure (Fig. 7), where the heat input was added for a total
LQR-FF 432.5
of four manipulated variables used to control three temperatures.
The MPC controller showed better performance for disturbances DSB 1000
not only in the feed flowrate but also in the feed composition. Nev- DSV 790
ertheless, the additional manipulated variable casts some doubt on
LSB 1000
the conclusion of Adrian et al. [42] that MPC alone is generally better
than a multi-SISO control (e.g. conventional multi-loops PID). LSV 714 minutes

4.6.2. Controlling the temperature in the bottom of the 0 200 400 600 800 1000
prefractionator
Wang and Wong [25] studied a high-purity ethanol/1- Settling time for +10% persistent disturbance in xA feed
propanol/1-butanol column. PID algorithms and temperature
μ-controller 569
control were used, instead of composition control. A temperature
in the prefractionator and two temperatures in the main column LSDP 569
were selected. The pairing was completely different than that pre- LQR-IA 839
viously proposed by Adrian et al. [42]. Fig. 6 (right) shows that a
temperature in the bottom section of the prefractionator was con- LQR-FF 0
trolled by manipulating reboiler heat duty while a temperature in DSB 1000
the rectifying section was controlled by manipulating the reflux
DSV 561
flow rate. Moreover, a temperature near the base of the column
was controlled by manipulating the side stream flow rate. Again, LSB 1000
the liquid split ratio (ˇL ) was not used for control. Stable control LSV 525 minutes
was achieved and products returned to their desired purity levels
for feed flow rate changes. However, large product purity devia- 0 200 400 600 800 1000
tions were reported for feed composition disturbances [25]. The
Fig. 10. Settling time for +10% disturbance in the feed flowrate (left) and feed com-
authors recommended a temperature/composition cascade control
position (right).
structure (Fig. 8) to solve this problem.
288 A.A. Kiss, C.S. Bildea / Chemical Engineering and Processing 50 (2011) 281–292

0.985 0.985
Disturbance +10% F
0.980 0.980
xB
Mole fraction / [-]

Mole fraction / [-]


0.975 0.975
xB
0.970 0.970
xA
0.965 0.965
xA
0.960 0.960 xC
0.955 0.955
xC
Disturbance +10% F
0.950 0.950
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Time / [hr] Time / [hr]

0.985 0.985
xA
0.980 xA 0.980
Mole fraction / [-]

Mole fraction / [-]


0.975 0.975
xC xB
0.970 0.970
xC
0.965 xB 0.965

0.960 0.960

0.955 0.955
Disturbance +10% xA Disturbance +10% xA
0.950 0.950
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Time / [hr] Time / [hr]

0.985 0.985

xB
Mole fraction / [-]

Mole fraction / [-]

0.980 0.980 xA
xA xC

0.975 0.975
xB
xC
0.970 0.970

SP +1% SP +1%
0.965 0.965
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Time / [hr] Time / [hr]

Fig. 11. Dynamic response of the LSV control structure, at a persistent disturbance Fig. 12. Dynamic response of the MPC control structure, at a persistent disturbance
of +10% in the feed flow rate (top), +10% xA in the feed composition (mid) and +1% of +10% in the feed flow rate (top), +10% xA in the feed composition (mid) and +1%
increase of the setpoint (bottom). increase of the setpoint (bottom).

4.7. Feedforward control to reject frequent measurable located in the main column and one temperature on the prefrac-
disturbances tionator side of the wall, were used to adjust the four manipulated
variables. Feed flow rate disturbances are well handled with this
Ling and Luyben [29] proposed to control the impurity levels structure, but product purities start to deviate significantly from
in the three product streams and one composition in the pre- their desired values for feed composition changes greater than 10%.
fractionator. This implicitly achieves also the minimization of the In a second proposed control structure, four differential tempera-
energy requirements. The four manipulated variables are liquid ture control loops were used. The performance is improved and
split, reflux flow rate, side stream flow rate, and vapor boilup. disturbances of 20% in feed composition are also well handled with
Fig. 9 shows the four-point control structure using combined only small deviations in product purities. This second structure also
feedback–feedforward to account for feed flow disturbances handles large changes in the column operating pressure [41].
In addition, the authors suggest combining feedback and feed-
forward to improve the performance with respect to feed flow rate 4.8. Advanced control techniques
disturbances. The improved performance was demonstrated by the
results of the dynamic simulation. Alstad and Skogestad [43] described the null space method
In a more recent paper [41], the authors explored also the use as a self-optimizing control method that selects the control vari-
of temperature measurements in order to avoid expensive and ables as combinations of measurements. For the case of a Petlyuk
high-maintenance composition analyzers. Two types of tempera- distillation setup this resulted into the following candidate mea-
ture control structures were studied. In the first, three temperatures surements: temperature at all stages and all flow rates. Using the
A.A. Kiss, C.S. Bildea / Chemical Engineering and Processing 50 (2011) 281–292 289

2
PC
Prefractionator V1 LC
L 1.8 V 2 / [kmol/h] = 56
L1

Q / [106 kcal/h]
D
1.6 62
FC TC
TC 1.4
F S 68
1.2
V2 TC CC 74
Input 1
LC select
20 25 30 35 40
L2 Main column
TC
L 1 / [kmol/h]
LC

B Fig. 14. Pentane–hexane–heptane separation in a DWC: energy requirements vs.


Q liquid flow rate for different values of the vapor flow rate (the dot marker shows the
operating point).
Fig. 13. Two-column model of a dividing wall distillation column.

Hereby we limit to present only the dynamic results for the best PID
null space method a subset of six measurements was obtained, control structure (LSV) and linear model predictive control (MPC).
resulting in a practically implementation. The PID control loops were tuned by the direct synthesis method
Woinaroschy and Isopescu [44] showed the ability of iterative [28]. As fairly accurate evaluations of the process time constants
dynamic programming to solve time optimal control of DWC. Their , 20 and 40 min were used. For the level controllers, a large reset
study focuses on startup control of the DWC. time  i = 100 min was chosen as no tight control is required. The liq-
These different studies investigate different separation systems uid levels in the reflux tank and reboiler are maintained by means
and types of disturbances and hence a common conclusion to iden- of D (distillate) and B (bottoms flow rate) whereas the product
tify the best controller cannot be drawn. In a comparative study, compositions are maintained by manipulating L (liquid reflux), S
van Diggelen et al. [45] applied more advanced controllers such (side product flowrate) and V (vapor boil-up). As illustrated by the
as LQG/LQR, GMC and higher order controllers based on H∞ norm next figures, the mole fractions of components benzene in the top
␮-synthesis. The performance was compared to several structures distillate (xA ), toluene in the side stream (xB ) and xylene in the bot-
involving multi-loop PID controllers. The controllers were applied tom product (xC ) are returning to their set point (SP = 0.97) within
to a DWC used in the ternary separation of benzene/toluene/xylene fairly short settling times. The control structure LSV exhibits rea-
(BTX). sonably low overshooting and short settling times, as demonstrated
Fig. 10 shows the settling times for +10% disturbances in the by Fig. 11.
feed flowrate and feed composition, respectively. While PI control Note that the dynamic DWC model used for testing the control
structures were also able to control the DWC, significantly shorter strategies is not a reduced one but a full-size nonlinear model that
settling times and better control were achieved using MIMO con- is representative of industrial applications. The quality of the lin-
trollers. The LQG with integral action and reference inputs was earized model used for the predictions inside MPC is derived from
found to deliver the best control performance. However, LQG was and tested against the nonlinear model. The MPC controller was
unable so far to make a significant industrial impact, mainly due to designed to handle a 6 × 6 system of inputs and outputs. The inputs
the fact that LQC cannot address well the constraints on the pro- include the controlled variables – mole fraction of benzene in dis-
cess inputs, states and outputs but also due to other limitations in tillate (xA ), toluene in the side stream (xB ), xylene in the bottom
handling the process nonlinearities and re-optimization algorithm product (xC ) and xylene in the top of the prefractionator (YC PF1 ),
required at every step [46]. Among the multi-loop PID strategies, as well as liquid holdups in the reflux tank (Ht ) and reboiler (Hr ).
LSV and DSV were the best, being able to handle persistent distur- The outputs include the manipulated variables – D, B, L, S, V and ˇL .
bances in reasonably-short times (Fig. 10). This agrees very well For the MPC tuning, a trial and error method was used as no reliable
with Abdul Mutalib and Smith [38], although Wolff and Skogestad design rules are available in the literature. Basically, the tuning of
[36] reported the lack of controllability for the LVS control structure MPC relies on a number of factors related to the controller as well as
– which was attributed to the nonlinearity of the response from V the process – prediction (p) and control (m) horizon, input (wu ) and
towards xS2 . output (wy ) weights, operating constraints on inputs and outputs
as well as change rate of inputs (u). The following key parameters
5. Case studies were used: p = 20, m = 3, and sampling time k = 3 min.
The dynamic response of the MPC controller under the oper-
5.1. Benzene–toluene–xylene separation ating constraints is shown in Fig. 12, being characterized by low
overshooting and short settling times. It is clear that MPC delivers
In order to allow a fair comparison with previously published an excellent performance in case of different industrially relevant
references, this work considers as a first case-study the separation disturbances and set point tracking. Compared to the open loop
of the ternary mixture benzene–toluene–xylene (BTX). The design simulation response of the process, the MPC reacts naturally upon
of the DWC and the dynamic model were described in detail by van the disturbances and it is able to steer the system to the given set
Diggelen et al. [45]. Fig. 2 illustrates the modeled DWC, consisting points under the specified constrains.
of 6 sections of 8 stages each. The task of the DWC is to split an
equimolar BTX mixture into three fractions of 97% purity each. 5.2. Pentane–hexane–heptane separation
Unlike the work of van Diggelen et al. [45], the next control struc-
tures make use of an extra loop controlling the heavy component This section presents results of design, dynamic simulation
in the top of the prefractionator, by manipulating the liquid split and control of a DWC for the separation of 100 kmol/h pen-
ratio and thus implicitly achieving energy minimization [39,40,29]. tane/hexane/heptane mixture 20:60:20 (molar). For the same
290 A.A. Kiss, C.S. Bildea / Chemical Engineering and Processing 50 (2011) 281–292

1 x B,3 1 x B,3
x D,1 x D,1

Purity / [-]

Purity / [-]
0.95 0.95

x S,2 x S,2
0.9 0.9

CS 1 CS 2
0.85 0.85
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Time / [h] Time / [h]
1 1

0.995 0.995
Purity / [-]

Purity / [-]
x D,1 x D,1
0.99 0.99
x B,3 x B,3 x S,2
0.985 x S,2 0.985
CS 3 CS 4
0.98 0.98
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Time / [h] Time / [h]

Fig. 15. Results of dynamic simulations for the separation of a pentane–hexane–heptane mixture in a DWC (control structures CS 1–4).

mixture, the controllability analysis of a Petlyuk arrangement [47] 1.074 × 106 kcal/h by adjusting the liquid and vapor flow rates
predicted easy rejection of disturbances, which was confirmed later from the main column to the prefractionator, L1 = 27.55 kmol/h and
by rigorous dynamic simulations. However, the system could not V2 = 68.51 kmol/h (Fig. 14). However, in this section we restrict to
cope with a large and fast change of the hexane feed flow rate. the sub-optimal design presented in Table 3.
Remarkably, this control problem was apparent during pressure A full pressure-driven dynamic simulation was built in Aspen
driven simulation, but could not be identified by linear analysis or Dynamics. Four control strategies denoted by CS1, CS2, CS3 and CS4
by flow-driven simulations in Aspen Dynamics. were investigated. In addition to the standard pressure and level
The preliminary design of the DWC – obtained by applying the control loops, temperatures at the top (tray 4) and bottom (tray 36)
short-cut method recommended by Triantafyllou and Smith [48] of the main column were controlled by the reflux and reboiler duty,
– was the starting point of rigorous simulations in Aspen Plus and respectively. The side-stream flow rate controlled a temperature
Aspen Dynamics. The distillate, reflux and side stream flow rates below the side-stream (tray 27) in CS1 and CS2 and the side-stream
were adjusted to meet the purity specifications. Weir heights of purity in CS3 and CS4. The flow rate of liquid going from the main
5 cm and residence times of 5 min in the reflux drum and 10 min in column to the prefractionator (L1 ) was kept constant in CS1 and
column sump were assumed. Condenser and tray pressure drops CS3 or used to control the temperature on the prefractionator top
of 0.05 bar and 0.01 bar, respectively, were specified. Pumps were (tray 3) in CS2 and CS4.
chosen for a 0.3 bar pressure drop over the corresponding valve. Fig. 15 presents the dynamic simulation results for the following
The complete flowsheet is presented in Fig. 13 (not all scenario. The simulation starts from the steady state. After 0.5 h,
pumps and valves are shown), while Table 3 summarizes the the feed rate is increased by 10%, from 8617 kg/h to 9400 kg/h.
results of sizing the columns. The energy requirement of the At time t = 3 h, the feed rate is reduced to 90% of the initial value
design was 1.164 × 106 kcal/h which could be further reduced to (7600 kg/h).
For all control structures, the composition of top and bottom
streams is almost constant, even if only temperature measure-
Table 3 ments are used. However, controlling a temperature in the middle
Design of Petlyuk distillation system.
section of the main column (CS1 and CS2) does not ensure the
Design parameter Value purity of the side stream product, which drops from 0.99 to 0.9,
Prefractionator for a 10% increase of the feed flow rate. Moreover, when the feed
Number of trays 12 rate is brought back to the initial value or even reduced to 90% of
Feed tray 5 it, the purity does not recover. This behavior suggests the existence
Diameter/(m) 0.8
of multiple steady states.
Top stage pressure/(bar) 1.17
Main column The performance of the system dramatically improves when
Number of trays 42 composition measurement is available and the side-stream flow
Top feed/withdraw tray 6 rate is used to control its purity (CS3 and CS4). In this case, the
Side stream product 15 purities of top and bottom products deviate by less than 0.3% from
Bottom feed/withdraw tray 32
Diameter/(m) 1.2
the design values. The performance could be further improved by
Top stage pressure/(bar) 1 adding composition–temperature cascade loops. Moreover, excel-
Distillate rate/(kmol/h) 19.77 lent control is achieved for the side-stream composition. Control of
Side stream rate/(kmol/h) 60.20 the prefractionator top temperature (CS4) improves the dynamics
Reflux rate/(kmol/h) 160.22
of the side-stream composition control loop. It should be remarked
Liquid to prefractionator, L1 /(kmol/h) 31
Vapor to prefractionator, V2 /kmol/h) 62 that, for all control structures investigated, the system becomes
Reboiler duty/(106 kcal/h) 1.164 unstable if, instead of liquid rate (L1 ), the liquid split ratio (ˇL ) is
Condenser duty/(106 kcal/h) 1.123 kept constant.
A.A. Kiss, C.S. Bildea / Chemical Engineering and Processing 50 (2011) 281–292 291

The controllability of the design with minimum energy require- [4] Z. Olujic, M. Jodecke, A. Shilkin, G. Schuch, B. Kaibel, Equipment improve-
ments was also investigated. This design appeared to be more ment trends in distillation, Chemical Engineering and Processing 48 (2009)
1089–1104.
sensitive to disturbances even if the purity of the side stream was [5] N. Asprion, G. Kaibel, Dividing wall columns: fundamentals and recent
controlled. For example, a 10% increase of the flow rate desta- advances, Chemical Engineering and Processing: Process Intensification 49
bilized the system if performed stepwise. This change could be (2010) 139–146.
[6] I. Dejanović, L. Matijašević, Ž. Olujić, Dividing wall column—a breakthrough
implemented only by ramping it over 6 h. towards sustainable distilling, Chemical Engineering and Processing: Process
Intensification 49 (2010) 559–580.
[7] F.B. Petlyuk, V.M. Platonov, D.M. Slavinskii, Thermodynamically optimal
6. Conclusions method for separating multicomponent mixtures, International Chemical Engi-
neering 5 (1965) 555–561.
Some broad guidelines and rules for the design of the DWC as [8] A.C. Christiansen, S. Skogestad, L. Liena, Complex distillation arrangements:
extending the Petlyuk ideas, Computers & Chemical Engineering 21 (1997)
well as standards for selecting control structures for DWC were
237–242.
recommended [49]. These were used in several studies, described [9] M.A. Schultz, D.G. Stewart, J.M. Harris, S.P. Rosenblum, M.S. Shakur, D.E. O’Brien,
previously. Hereby, the following guidelines can be recommended Reduce costs with dividing-wall columns, Chemical Engineering Progress 2002
(May) (2002) 64–71.
in order to select the appropriate control strategy for DWC. Note
[10] B. Kolbe, S. Wenzel, Novel distillation concepts using one-shell columns, Chem-
that they correspond to the history of developing control structures ical Engineering & Processing 43 (2004) 339–346.
for the dividing-wall distillation systems. [11] H. Becker, S. Godorr, H. Kreis, Partitioned distillation columns—why, when &
how, Journal of Chemical Engineering (January) (2001) 68–74.
[12] B. Kaibel, H. Jansen, E. Zich, Z. Olujic, Unfixed dividing wall technology for
• Use the reflux and boilup streams to control the distillate and packed and tray distillation columns, Distillation and Absorption 152 (2006)
bottom compositions. 252–266.
• Alternatively, use the distillate and bottoms flow rates for com- [13] R. Isopescu, A. Woinaroschy, L. Draghiciu, Energy reduction in a divided wall
distillation column, Revista de Chimie 59 (2008) 812–815.
position control. [14] A.A. Kiss, H. Pragt, C. van Strien, Reactive dividing-wall columns—how to get
• Use the side stream flow rate to control the side stream compo- more with less resources? Chemical Engineering Communications 196 (2009)
1366–1374.
sition.
[15] S. Midori, S.N. Zheng, I. Yamada, Azeotropic distillation process with vertical
• Alternatively, use the distillate and bottoms flow rates for com- divided-wall column, Kagaku Kogaku Ronbunshu 27 (2001) 756–760.
position control. [16] C. Bravo-Bravo, J.G. Segovia-Hernández, C. Gutiérrez-Antonio, A.L. Duran, A.
• Use the liquid split to control the amount of heavy component Bonilla-Petriciolet, A. Briones-Ramírez, Extractive dividing wall column: design
and optimization, Industrial & Engineering Chemistry Research 49 (2010)
leaving in the top of the prefractionator – this preserves the opti- 3672–3688.
mality characteristics of a given design. [17] I. Mueller, E.Y. Kenig, Reactive distillation in a dividing wall column—rate-
• Use (inferential) temperature measurements to avoid composi- based modeling and simulation, Industrial & Engineering Chemistry Research
46 (2007) 3709–3719.
tion measurements, combined when required with feedforward [18] S. Hernandez, R. Sandoval-Vergara, F.O. Barroso-Munoz, R. Murrieta-Duenasa,
ratio schemes. H. Hernandez-Escoto, J.G. Segovia-Hernandez, V. Rico-Ramirez, Reactive divid-
• Use concentration – temperature – cascade loops to improve dis- ing wall distillation columns: simulation and implementation in a pilot plant,
Chemical Engineering & Processing 48 (2009) 250–258.
turbance rejection. [19] P.B. Shah, Squeeze more out of complex columns, Chemical Engineering
• Combine conventional feedback control with feedforward con- Progress (July) (2002) 46–55.
trol loops. [20] M. Serra, A. Espuña, L. Puigjaner, Control and optimization of the divided wall
• Apply model predictive control (MPC) and/or other MIMO control column, Chemical Engineering and Processing 38 (1999) 549–562.
[21] M. Serra, A. Espuña, L. Puigjaner, Study of the divided wall column controlla-
strategies, when SISO loops are not sufficient. bility: influence of design and operation, Computers & Chemical Engineering
24 (2000) 901–907.
[22] M. Serra, M. Perrier, A. Espuña, L. Puigjaner, Analysis of different control pos-
This study indicates that the dividing-wall distillation columns sibilities for the divided wall column: feedback diagonal and dynamic matrix
have good controllability properties, provided that an appropriate control, Computers & Chemical Engineering 25 (2001) 859–866.
control structure is implemented. Among the multi-loop PID con- [23] M. Serra, A. Espuña, L. Puigjaner, Controllability of different multicomponent
distillation arrangements, Industrial and Engineering Chemistry Research 42
trol strategies, LSV and DSV were the best, being able to handle (2003) 1773–1782.
persistent disturbances in reasonably-short times. However, sig- [24] J.G. Segovia-Hernandez, E.A. Hernandez-Vargas, J.A. Marquez-Munoz, Control
nificantly shorter settling times and better control performance properties of thermally coupled distillation sequences for different operating
conditions, Computers & Chemical Engineering 31 (2007) 867–874.
were achieved using advanced multi-input multi-output (MIMO) [25] S.J. Wang, D.S.H. Wong, Controllability and energy efficiency of a high-purity
controllers such as model predictive control (MPC). divided wall column, Chemical Engineering Science 62 (2007) 1010–1025.
[26] A.A. Kiss, C.S. Bildea, A.C. Dimian, Design and control of recycle systems by
non-linear analysis, Computers & Chemical Engineering 31 (2007) 601–611.
Acknowledgements [27] J. Harmsen, Process intensification in the petrochemicals industry: drivers and
hurdles for commercial implementation, Chemical Engineering and Processing
49 (2010) 70–73.
We thank Zarco Olujic (TU Delft, NL), Igor Dejanovic (University
[28] W.L. Luyben, M.L. Luyben, Essentials of Process Control, McGraw-Hill,
of Zagreb, HR), Ivar J. Halvorsen (SINTEF, NO) and Sigurd Skoges- 1997.
tad (Norwegian University of Science and Technology) for the very [29] H. Ling, W.L. Luyben, New control structure for divided-wall columns, Industrial
helpful discussions on the operation and control of DWC. The sup- & Engineering Chemistry Research 48 (2009) 6034–6049.
[30] J.G. Segovia-Hernandez, S. Hernandez, A. Jimenez, Analysis of dynamic prop-
port given by Rohit Rewagad (University of Twente, NL) is also erties of alternative sequences to the Petlyuk column, Computers & Chemical
gratefully acknowledged. Engineering 29 (2005) 1389–1399.
[31] S. Robles-Zapiain, J.G. Segovia-Hernandez, A. Bonilla-Petriciolet, R. Maya-
Yescas, Energy-efficient complex distillation sequences: control properties,
References Canadian Journal of Chemical Engineering 86 (2008) 249–259.
[32] V.E. Tamayo-Galvan, J.G. Segovia-Hernandez, S. Hernandez, J. Cabrera-Ruiz,
[1] G. Kaibel, Distillation columns with vertical partitions, Chemical Engineering J.R. Alcantara-Avila, Controllability analysis of alternate schemes to complex
Technology 10 (1987) 92–98. column arrangements with thermal coupling for the separation of ternary
[2] Z. Olujic, B. Kaibel, H. Jansen, T. Rietfort, E. Zich, G. Frey, Distillation column mixtures, Computers & Chemical Engineering 32 (2008) 3057–3066.
internals/configurations for process intensification, Chemical & Biochemical [33] B.G. Rong, I. Turunen, A new method for synthesis of thermodynamically equiv-
Engineering Quarterly 17 (2003) 301–309. alent structures for Petlyuk arrangements, Chemical Engineering Research &
[3] Z. Olujic, L. Sun, M. Gadalla, A. de Rijke, P.J. Jansens, Enhancing ther- Design 84 (2006) 1095–1116.
modynamic efficiency of energy intensive distillation columns via internal [34] F.I. Gomez-Castro, J.G. Segovia-Hernandez, S. Hernandez, C. Gutierrez-Antonio,
heat integration, Chemical & Biochemical Engineering Quarterly 22 (2008) A. Briones-Ramirez, Dividing wall distillation columns: optimization and con-
383–392. trol properties, Chemical Engineering & Technology 31 (2008) 1246–1260.
292 A.A. Kiss, C.S. Bildea / Chemical Engineering and Processing 50 (2011) 281–292

[35] J.R. Alcantara-Avila, J. Cabrera-Ruiz, J.G. Segovia-Hernandez, S. Hernandez, B.G. [44] A. Woinaroschy, R. Isopescu, Time-optimal control of dividing-wall dis-
Rong, Controllability analysis of thermodynamically equivalent thermally cou- tillation columns, Industrial & Engineering Chemistry Research 49 (2010)
pled arrangements for quaternary distillations, Chemical Engineering Research 9195–9208.
& Design 86 (2008) 23–37. [45] R.C. Van Diggelen, A.A. Kiss, A.W. Heemink, Comparison of control strategies for
[36] E.A. Wolff, S. Skogestad, Operation of integrated three-product (Petlyuk) dis- dividing-wall columns, Industrial & Engineering Chemistry Research 49 (2010)
tillation columns, Industrial & Engineering Chemistry Research 34 (1995) 288–307.
2094–2103. [46] S.J. Qin, T.A. Badgwell, A survey of industrial model predictive technology, Con-
[37] P. Lundström, J.H. Lee, M. Morari, S. Skogestad, Limitations of dynamic matrix trol Engineering Practice 11 (2003) 733–764.
control, Computers & Chemical Engineering 19 (1995) 409–421. [47] C.S. Bildea, On the controllability of Petlyuk distillation systems, NPT Process
[38] M.I. Abdul Mutalib, R. Smith, Operation and control of dividing wall distillation Technologies 11 (2004) 21–25.
columns. 1. Degrees of freedom and dynamic simulation, Transactions of the [48] C. Triantafyllou, R. Smith, The design and optimisation of fully ther-
Institution of Chemical Engineers 76 (Pt. A) (1998) 308–318. mally coupled distillation columns, Transactions IChemE 70 (Pt. A) (1992)
[39] I.J. Halvorsen, S. Skogestad, Optimizing control of Petlyuk distillation: under- 118–132.
standing the steady-state behaviour, Computers & Chemical Engineering 21 [49] K. Il Kim, M. Lee, S. Park, Dynamic simulation for the structural design of the
(1997) 249–254. divided wall column for different feed composition and various separation fea-
[40] I.J. Halvorsen, S. Skogestad, Optimal operation of Petlyuk distillation: steady- tures, in: International Conference on Control, Automation and Systems, vol.
state behavior, Journal of Process Control 9 (1999) 407–424. 1–6, 2007, pp. 1988–1992.
[41] H. Ling, W.L. Luyben, Temperature control of the BTX divided-wall column, [50] M.I. Abdul Mutalib, A.O. Zeglam, R. Smith, Operation and control of dividing wall
Industrial & Engineering Chemistry Research 49 (2010) 189–203. distillation columns. 2. Simulation and pilot plant studies using temperature
[42] R. Adrian, H. Schoenmakers, M. Boll, MPC of integrated unit operations: control control, Transactions of the Institution of Chemical Engineers 76 (Pt. A) (1998)
of a DWC, Chemical Engineering & Processing 43 (2004) 347–355. 319–334.
[43] V. Alstad, S. Skogestad, Null space method for selecting optimal measure-
ment combinations as controlled variables, Industrial & Engineering Chemistry
Research 46 (2007) 846–853.

You might also like