You are on page 1of 14

Chemical Engineering and Processing 43 (2004) 421–434

Model-based design, control and optimisation of


catalytic distillation processes
C. Noeres, K. Dadhe, R. Gesthuisen, S. Engell, A. Górak∗
Lehrstuhl für Thermische Verfahrenstechnik, Fachbereich Bio-und Chemieingenieurwesen, Universität Dortmund, D-44221 Dortmund, Germany

Received 18 December 2002; received in revised form 15 May 2003; accepted 20 May 2003

Abstract

In this contribution, the benefits of using dynamic models of different complexity and size for process design, optimal operation and control
of catalytic distillation (CD) processes are discussed for the case study of the heterogeneously catalysed reactive distillation (RD) of methyl
acetate. Dynamic reactive distillation experiments at pilot plant scale were performed using the catalytic structured packing MULTIPAK. A
dynamic rate-based model were developed which contains hydrodynamic effects as liquid holdup, liquid backmixing and pressure drop as
well as reaction kinetics and which describes the process behaviour accurately. For offline and online optimisation and control, reduced order
and simplified models were applied. A systematic control structure selection and controller design studies for the experimental column were
conducted. The linear controller shows good performance over a wide range of operating conditions.
© 2003 Elsevier B.V. All rights reserved.

Keywords: Reactive distillation; Catalytic distillation; Catalytic packing; Methyl acetate; Modelling; Control

1. Introduction KATAPAK-SP [5], catalytic beads are immobilised within


gauze pockets. The alternation of two different layers (“cat-
Reactive distillation (RD), the combination of chemical alytic pockets” and “open channels”) results to a so-called
reaction and distillation in a single column, is one of the hybrid sandwich structure which allows easy access of the
most important industrial applications of the multifunc- liquid to the catalyst and free paths for the rising vapour
tional reactor concept. Recently, it has drawn considerable to ensure counter-current operation in presence of small
attention because of its striking advantages, especially catalyst particles without flooding problems. In this way,
for equilibrium-limited reactions. Several reviews have structured catalytic packings permit excellent solid–liquid
been published in the last decade which give an excellent as well as liquid–vapour mass transfer.
overview of RD processes [1–3]. Reactive distillation pro- In catalytic distillation, thermodynamic and diffusional
cesses can be divided into homogeneous processes, either couplings in the phases and at the interface which are of
auto-catalysed or homogeneously catalysed, and hetero- multicomponent character are accompanied by complex
geneous processes, in which the reaction is catalysed by chemical reactions. As a consequence, special rate-based
a solid catalyst. The latter, often referred to as catalytic mathematical models which take the column hydrodynam-
distillation (CD), permits an optimum configuration of the ics, mass transfer resistances, reaction kinetics and dynamic
reaction and separation zones in a RD column and the process behaviour into account are required to describe such
expensive recovery of liquid catalysts can be avoided. processes adequately for process design and scale-up issues.
The most promising column internals for CD are Steady-state experiments for the synthesis of methyl ac-
structured catalytic packings that combine favourable char- etate [6] and various ethers like MTBE and TAME [7] have
acteristics of traditional structured packings and heteroge- already been published. However, experimental data for the
neous catalysts. In these packings, like MULTIPAK [4] or analysis of the dynamic behaviour of RD columns with struc-
tured catalytic packings are rarely available.
∗ Corresponding author.
The availability of a reliable dynamic model is the pre-
E-mail address: a.gorak@tv.chemietechnik.uni-dortmund.de condition for model-based process control and optimisation.
(A. Górak). On the one hand the integration of reaction and separation

0255-2701/$ – see front matter © 2003 Elsevier B.V. All rights reserved.
doi:10.1016/j.cep.2003.05.001
422 C. Noeres et al. / Chemical Engineering and Processing 43 (2004) 421–434

is advantageous from the design point of view as the same 2. Rigorous modelling
task is performed with less equipment and less energy, but
on the other hand it creates a challenge to control. Com- Extensive surveys on modelling approaches for catalytic
pared to reaction and separation processes in series, the distillation processes have been published in [2,3]. How-
integration leads to a reduction of the number of manipu- ever, only a limited number of publications deal with the
lated variables, measurements and also of operational de- transient behaviour of CD columns equipped with catalytic
grees of freedom [8]. Therefore tight control is required as structured packings. Most contributions are based on equi-
disturbances may lead to a significant loss of performance. librium models assuming thermodynamic and chemical
For the design of an appropriate control structure and a equilibrium on each stage. However, equilibrium is rarely
controller, model-based approaches can be applied. Control attained since mass and heat transfer as well as chem-
structure selection for RD processes was carried out previ- ical reactions are rate driven by chemical potential and
ously by open-loop and closed-loop simulations of different temperature.
scenarios or by investigations of the non-linear steady-state Kreul published a dynamic rate-based model, mod-
behaviour [9]. For a systematic control structure selection elling vapour and liquid phase separately, for packed RD
and also for the optimisation of batch trajectories simplified columns [10]. The reaction kinetics were implemented
models have to be applied. These simplified models can be by means of an overall kinetic approach lumping sev-
derived from complex models. eral effects as reaction kinetics, liquid backmixing and
The aim of this paper is to give a comprehensive overview mass transfer resistances together. Further investigations
of the use of dynamic modelling for different tasks, e.g. dy- have shown, that this approach yields good predictions in
namic simulation, scale-up, controller design and optimisa- a certain operating range but leads to deviations if it is
tion, in the synthesis of methyl acetate in a CD column. The extrapolated to different process conditions. In this contri-
derived models are validated with experimental data from a bution, the approach has been extended by including the
pilot plant. The rigorous model is used as a reference model catalyst phase as well as considering the non-ideal flow
for simulations of the closed-loop performance. behaviour in the liquid phase. This is essential to reflect the
The paper is divided into four chapters: in the first one we complex hydrodynamics occurring in structured catalytic
discuss the rigorous model development, necessary model packings.
parameters as mass transfer coefficients, hydrodynamic
quantities and reaction kinetics. Then the experimental 2.1. Modelling approach
set-up of the used RD pilot plant is presented. Different ex-
perimental runs are used for the model validation and shown The rate-based-model (RBA) used in this paper is based
in Section 3. Afterwards, optimisation issues are discussed on the two-film theory and comprises the material and energy
and a systematic control structure selection and controller balances of a differential element of the vapour and of the
design studies for the experimental column are presented. liquid phase (see Fig. 1). The dynamic component balances

B B B B
L,x, i T L V,y, i TV
z z

δ Dax,L δ δ
S/L F F V/L V/L
z F,x , T
L i L L
Dax,V

NLi
NSi B
QS HL
T V
S UL,t Q L
QL
T L B
S T L
T yIi
I
y
B
x i
i
c/z
I
x
B
xi
∆H R
0
R i
HL
UV,t Q V

Dax,L Dax,V
B B
L,x, T z+dz V,y B
i,
B
i L V
z+dz

S/L V/L V/L


Catalyst Liquid Liquid bulk phase Liquid Vapour Vapour bulk phase
film film film

Fig. 1. Three-phase rate-based approach for catalytic distillation processes.


C. Noeres et al. / Chemical Engineering and Processing 43 (2004) 421–434 423

for the liquid and the vapour are: information about the hydrodynamics of the columns inter-
nals as structured packings and liquid distributors as well as
∂ULi Dax,L ∂2 ∂ the column periphery such as the top and bottom section.
= (LxB
i )− (LxB )
∂t uL ∂z2 ∂z i Liquid holdup and pressure drop correlations were derived
+(NLi ai + NSi acat )Ac ; i = 1, . . . , nc (1) for the catalytic packing MULTIPAK [10]. Liquid backmix-
ing in MULTIPAK was extensively studied by Noeres et al.
∂ i
0= i ) − NLi a Ac .
(VyB (2) [15].
∂z Correlations for the description of the liquid holdup and
The dynamic energy balances are: pressure drop in the non-catalytic packing ROMBOPAK-6M
were published by Mackowiak [16]. The hydrodynamic
∂EL αax,L ∂2 ∂ characteristics of the column periphery were determined at
= (Lh̄L ) − (Lh̄L )
∂t uL ∂z2 ∂z pilot scale (see Section 3).
+(QL ai + QS acat )Ac − QHL
L ; i = 1, . . . , nc
(3) 2.4. Chemical system and reaction kinetics

0= (V h̄V ) − QL ai Ac − QHL
V . (4) The synthesis of methyl acetate from methanol and acetic
∂z acid (HAc) is a slightly exothermic equilibrium limited liq-
Because the column is operated at atmospheric pressure the uid phase reaction:
vapour holdup is negligible. Moreover, reaction rates are
taken into account by considering the solid catalyst phase. CH3 OH + (CH3 )COOH ↔ (CH3 )COO(CH3 ) + H2 O
Due to the rate limited kinetics, diffusion phenomena inside
the catalyst are neglected. HR0 = −4.2 kJ/mol (9)

2.2. Mass and heat transfer The low equilibrium constant and the strongly non-ideal
behaviour which causes the forming of the binary azeotropes
Diffusional interaction phenomena like osmotic or reverse methyl acetate/methanol and methyl acetate/water make this
diffusion which may occur in multicomponent systems are reaction system interesting as a possible RD application [17].
taken into account via the Maxwell–Stefan equations for Therefore, the methyl acetate synthesis was chosen as a test
mass transfer (see, e.g. [11,12]): system. Since the process is carried out under atmospheric
pressure the maximum process temperature is lower than
Ni = Ji + Nt χi ; i = 1, . . . , nc − 1. (5) 390 K and no side reactions in the liquid phase like the
The diffusive fluxes in the regarding film are calculated by: formation of diethyl ether occur [18].
Several authors published different approaches to describe
(J) = −ct [R]−1 [Γ ](χ). (6) the heterogeneously catalysed reaction for this chemical sys-
The elements of matrix [R] are defined as: tem. Kreul identified an overall kinetic by trickle-bed reactor
experiments lumping intrinsic kinetics, mass transfer resis-
χi 
nc
χk tances as well sorption effects together [10]. Pöpken et al.
Rii = + (7)
βinc βik carried out an extensive study on the methyl acetate syn-
k=1k=i
thesis using Amberlyst 15W [19]. Own laboratory experi-
 
1 1 ments were conducted to investigate the reaction kinetics of
Rij = −χi − . (8) the methyl acetate reaction catalysed by Lewatit K2621. A
βij βinc
modified LHHW-model published by Poepken was refitted
The binary mass transfer coefficient βij may be calculated using our new reaction kinetics data as well as binary ad-
using a Sherwood correlation for binary systems. It was pub- sorption data:
lished elsewhere for the catalytic structured packing MUL-
TIPAK [6]. In this study, mass transfer resistances between 

1 dni k1 aHAc aMeOH − k−1 aMeAc aH 2O
the liquid bulk phase and the catalyst pellets are taken into r= = mcat
consideration according to Wakao and Kaguei [13]. The νi dt (aHAc + aMeOH + aMeAc + aH 2O
)2
mass transfer in the non-catalytic section equipped with (10)
ROMBOPAK-6M was calculated using the correlations pub-
lished by Pelkonen [14]. Ki ai
ai = . (11)
Mi
2.3. Hydrodynamics
The temperature dependence of the reaction rate constants
Modelling approaches for the simulation of the transient are expressed by the Arrhenius equation. All reaction kinetic
behaviour of catalytic distillation columns require detailed parameters are given by Table 1.
424 C. Noeres et al. / Chemical Engineering and Processing 43 (2004) 421–434

Table 1 Table 3
Parameter set for reaction kinetics model (Eqs. (10) and (11)) Column configuration and initial process conditions
Adsorption-based approach Run A Run B
EA,1 (J/mol) 5.94 × 104
EA,−1 (J/mol) 6.09 × 104 Column top pressure, p (bar) 1.013 1.013
Initial methanol holdup, 1.248 1.140
k10 (mol/(g s)) 8.00 × 106
UMeOH (mol)
k−1
0
(mol/(g s)) 6.18 × 105
Feed position, hfeed (m) 2 2
σ̄ (%) 4.51
Column internals (2–3 m) ROMBOPAK-6M ROMBOPAK-6M
Binary adsorption data Column internals (1–2 m) MULTIPAK-I MULTIPAK-I
KHAc 0.24 Column internals (0–1 m) MULTIPAK-I MULTIPAK-I
KMeOH 1.57
KMeAc 0.125
KH 2 O 1.60
ms /mcat 0.56 plete dynamic behaviour of the system as well as changes
of the total and specific component molar holdups in every
segment are considered.
2.5. Thermodynamics and physical properties

The vapour–liquid equilibrium was calculated using the 3. Experimental set-up


modified Wilson equation [10]. For the vapour phase, the
dimerisation of acetic acid was taken into account using the All experiments were performed in the pilot plant shown
chemical theory to correct vapour phase fugacity coefficients in Fig. 2. The glass column has a inner diameter of 100 mm
[20]. Binary diffusion coefficients for the vapour phase and and an effective packing height of 3 m. The operating
for the liquid phase were estimated via the methods pur- pressure lies between 10 and 1013 mbar and the operating
posed by Fuller et al. and Tyn and Calus, respectively (see temperature is up to 390 K. The column is equipped with
[21]). Physical properties as densities, viscosities and ther- catalytic (MULTIPAK-I) and non-catalytic (ROMBOPAK-
mal conductivities were calculated by the methods given in 6M) structured packings. A detailed description of the pack-
[21]. Heat losses through the column wall were measured at ings and of the column configuration is given in Tables 2
pilot scale. and 3.
Feed and product streams, the heat duty and the column
2.6. Numerical solution wall heating jackets are controlled by a process control sys-
tem. All process data like temperatures, flow rates, liquid
The rigorous model presented above yields to a system of levels and system pressures are collected and monitored by a
partial differential and algebraic equations. A discretisation process control system. Liquid phase samples are withdrawn
in axial direction (column height) results in a complex and from the three liquid distributors (Q2, Q6, Q10) as well as
highly non-linear differential-algebraic equation (DAE) sys- from the reboiler (Q15). The samples are analysed by a gas
tem, which can only be solved numerically. Therefore, all chromatograph including a flame ionisation detector (FID)
model equations have been implemented into the dynamic and a thermal conductivity detector (TCD). Moreover, the
modelling environment gPROMS [22]. liquid phase composition in the reflux line (QI 501) and in
For the numerical solution, a careful analysis of the the reboiler (QI 502) are determined by an online near in-
full system of equations and determination of appropriate frared spectrometer (NIR).
initial conditions was necessary in order to prevent high The experimental design was supported by theoretical pa-
index problems. As these problems often arise in process rameter studies to determine an appropriate operating range
engineering applications due to simplifications that impose for the experimental test runs to investigate the dynamic pro-
additional constraints on the differential variables, the com- cess behaviour.

Table 2
Characteristic geometrical data of the catalytic and non-catalytic packings used in this work.
Geometrical data MULTIPAK-I (catalytic) ROMBOPAK-6M (non-catalytic)

Height of packing segment, Hp (mm) 166 130


Number of packing segments, n 6 8
Catalyst volume faction, ψcat 0.256 –
Specific surface area, a (m2 /m3 ) 370 230
Void fraction, ε 0.654 0.95
Inclination angle, θ (◦ ) 60 –
Effective catalyst beads size, dp (mm) 0.50–0.62 –
Corrugation spacing, S (mm) 7.5 –
C. Noeres et al. / Chemical Engineering and Processing 43 (2004) 421–434 425

Z1

Z1

Z1 Z0

Z0 Z0

Z0

Z0
Z0

Z0

Z0

Z0

Z1
Z0

Z0

N2

Z1 Z0

Z0

Z2 Z0 Z0

Fig. 2. Flow sheet of RD pilot plant column.


426 C. Noeres et al. / Chemical Engineering and Processing 43 (2004) 421–434

Fig. 3. Steady-state methyl acetate concentration profiles in dependence upon the modified reflux ratio RT as well as the heat duty Q.

4. Experimental model validation All test runs presented in this contribution were performed
as follows: The column was heated up with the initial amount
Fig. 3 shows the simulated steady-state methyl acetate of methanol given in Table 3 under total reflux conditions
concentration profiles for different operating conditions. without acetic acid feed. After steady-state conditions were
The distillate composition in the optimum operating range reached, the acetic acid was fed into the column (t = 0 s).
shows only a minor dependency upon the reflux ratio. For The complete schedule of the experimental runs is given in
model validation it is necessary to measure the column pro- Table 4.
files as well. Several test runs were conducted to investigate Figs. 4 and 5 show the predicted and the measured liquid
the transient process behaviour and to collect dynamic data concentrations and temperatures at different heights along
for a sufficient model validation. the column. At the beginning of run A the acetic acid feed

Table 4
Experimental schedule of dynamic RD experiments
Run A Run B
Ia IIa IIIa Ia IIa IIIa
Duration, t (s) 4680 5580 10950 1800 7590 9600
Heat duty, Q (W) 6000 6000 6000 6000 6000 6000
Reflux ratio, ν ∞ 10 1 ∞ 3 4
Feed rate, FHAV,2 m (mol/s) 0.024 0.024 0.024 0.027 0.027 0.027
a Period number.
C. Noeres et al. / Chemical Engineering and Processing 43 (2004) 421–434 427

Fig. 4. Dynamic liquid bulk concentrations of run A: experimental and simulation results (( ) acetic acid, (∗) methanol, (䊊) methyl acetate, (䊐) water).

Fig. 5. Dynamic liquid bulk temperatures of run A: experimental (- - -) and simulation results (—).
428 C. Noeres et al. / Chemical Engineering and Processing 43 (2004) 421–434

Fig. 6. Dynamic liquid bulk concentrations of run B: experimental and simulation results (( ) acetic acid, (∗) methanol, (䊊) methyl acetate, (䊐) water).

Fig. 7. Dynamic liquid bulk temperatures of run B: experimental (- - -) and simulation results (—).
C. Noeres et al. / Chemical Engineering and Processing 43 (2004) 421–434 429

is switched on and the methyl acetate concentration rises due to lower boiling trace components which are not con-
rapidly. At the end of period I the reflux ratio is changed sidered in the column model. After the start-up phase the
and distillate—a mixture of methyl acetate and methanol temperature along the catalytic section of the column shows
near the binary azeotrope—is withdrawn at the top of the a steep rise caused by the acetic acid feed. After a temper-
column. At the end of period II quasi-steady-state conditions ature maximum, the produced methyl acetate accumulates
are reached. The system is perturbed by a further change in inside the column and therefore the liquid temperature falls.
the reflux ratio and moves to a different operating point. In period II, distillate is withdrawn and in the catalytic sec-
The transient and the quasi-steady-state behaviour is pre- tion the low boiling component methyl acetate impoverishs.
dicted quite well by our rigorous model (see Figs. 4 and 5). The simulated temperature profiles are in good agreement
However, after the second perturbation (period III) the devi- with the experimental data.
ations between experiment and simulation are getting larger. The results presented here demonstrate clearly that our
Possible explanations for this could be on the one hand the rigorous model predicts the dynamic and quasi-steady-state
extreme operating conditions. Due to the low reflux ratio the behaviour of this RD process well.
liquid load drops from an average value of 5 m3 /(m2 /h) to
a value below 1.5 m3 /(m2 /h). This liquid load is extensively
below the load point [23] at the lower end of the loading 5. Control and optimisation
range of the catalyst packing which results in decreased cat-
alyst wetting and reduced reaction rate. As industrial RD Because of its complexity the rate-based model is not suit-
processes should be operated at higher liquid loads which able for controller design and optimisation of the RD pro-
ensure a sufficient catalyst wetting the model is able to pre- cess. Therefore an extended equilibrium stage model which
dict a broad range of reasonable process states with suffi- includes a reaction kinetic is used for these tasks. Fig. 8
cient accuracy. On the other hand, the deviations between shows a comparison of simulation results of the RBA and
experiment and simulation at the end of the test run can be the equilibrium stage model for a typical trajectory of input
explained by the batch character of all runs—small devia- variables.
tions between experiment and model are integrated by the The dynamic behaviour is covered well by the simplified
column reboiler over the process time. model and the deviations between the absolute values are
To examine the extrapolation capability of the dynamic acceptable for control purposes. The computational time for
model, further experiments at different operating conditions the simulation of the equilibrium stage model compared to
were performed. Figs. 6 and 7 show the transient liquid con- the RBA model is approximately 20 times smaller. This ad-
centration and temperature profiles of run B. Deviations of vantage motivates the use of the simplified model for control
the simulated liquid bulk temperature after the start-up are and optimisation purposes.

Fig. 8. Comparison of the RBA and the equilibrium stage model.


430 C. Noeres et al. / Chemical Engineering and Processing 43 (2004) 421–434

Table 5 Table 6
Available measurements and location in the column Remaining control structures after analysis of rhp zeros
Number Measurement Part of column Manipulated variables: Manipulated variables:
heat duty/reflux ratio HAc-feed/reflux ratio
1 Mole fraction HAc Condenser/reflux line
2 Mole fraction MeOH Condenser/reflux line Structures Smallest Structures Smallest
3 Mole fraction MeAc Condenser/reflux line rhp zero rhp zero
4 Mole fraction H2 O Condenser/reflux line 3 21 – 3 13 37.7811
5 Mole fraction HAc Reboiler 3 22 – 3 14 1.2408
6 Mole fraction MeOH Reboiler 3 9 0.1532 3 15 0.4353
7 Mole fraction MeAc Reboiler 3 10 0.0224 3 17 0.1340
8 Mole fraction H2 O Reboiler 3 11 0.0169 3 5 0.1257
9–12 Temperature Separation packing 3 8 0.0124 3 6 0.1165
13–16 Temperature Catalytic packing 1 3 6 0.0115 3 1 0.1029
17–20 Temperature Catalytic packing 2 3 18 0.0964
21 Temperature Reboiler 3 4 0.0930
22 Molar flow rate Condenser 3 9 0.0920
3 19 0.0701
3 12 0.0444
3 10 0.0403
5.1. Linear controller design 3 11 0.0334

The standard approach to control chemical processes is the


use of linear multi input multi output (MIMO) controllers. For the systematic control structure selection it is assumed
Due to the instationary character of the semi-batch process that the mole fraction of methyl acetate in the reflux is al-
at hand, the change in process dynamics has to be taken into ways chosen as control variable as it describes the product
account, i.e. a robust controller has to be developed. Dis- quality. Furthermore the ratio between the heat duty and
tillation processes usually offer a large number of possible the HAc-feed represents one degree of freedom. As conse-
control variables, e.g. temperatures, concentrations. For the quence it is impossible to use both variables as independent
semi-batch RD, available manipulated variables are the feed control inputs. Therefore we investigate systems with two
rate of acetic acid, the heat duty in the reboiler and the re- input and two output variables which yields in 42 potential
flux ratio. Potential control variables are the temperatures at control structures.
different locations and the product concentrations in the re- In a first selection step, the plant dynamics of all possi-
boiler and at the top of the column. Table 5 summarises the ble structures are checked for non-minimum phase elements
available measurements and their numbering. as the attainable performance of linear control systems is
For the 22 measurements and 3 manipulated variables limited by right-half-plane (rhp) zeros, time-delays and un-
numerous control structures can be chosen. In [24] a sys- stable poles [26]. By considering the multivariable zeros
tematic approach for control structure selection is proposed. and only the smallest delay, the attainable performance is
The described integrated approach yields a promising con- over-estimated. For the closed-loop a rise time of 500 s is
trol structure and a low order controller without a priori assumed. As the bandwidth is limited by 0.3ωN structures
assumptions on the controller structure and without a need with rhp zeros below 0.006 are excluded. The remaining
for many intuitive decisions by the designer. Starting from structures are given in Table 6.
a linear model of the given system with all possible inputs For further reduction of the number of combinations, the
and outputs, the procedure to determine the optimal control directionality of the remaining control structures is regarded.
structure applied to the RD is as outlined below. The directionality of the plant can be measured by the min-
In order to develop the linear controller for the semi-batch imised condition number γ ∗ . A large minimised condition
process in a first step a nominal trajectory has to be deter- number indicates that the plant gain varies strongly with
mined. Based on the equilibrium stage model Fernholz et al. the direction of the inputs, the gain matrix is then called
calculated an optimal trajectory for the manipulated vari- ill-conditioned. The advantage of using the minimised con-
ables by maximising the productivity of the process [25]. dition number is its scaling independence. A large min-
At a quasi-stationary point in the middle of this batch, imised condition number implies numerical problems in the
which is taken as the nominal operating point, a linear controller design and robustness problems if decoupling is
state space model is determined. The resulting linear sys- enforced. However, it is known that the condition number
tem, which was determined by linearisation of the non-linear is only critical around the closed-loop bandwidth which is
equilibrium stage model, has 85 states (the nominal values not known at this stage. So the condition numbers will only
of the manipulated variables RTs = 0.597, Qs = 3400 W, eliminate such structures which have a comparatively large
Fs = 0.04 mol/s). Under these conditions a methyl acetate condition number over a large frequency range. Only the
mole fraction of about 80% results. The dynamic behaviour structures {3 5}, {3 6} and {3 19} with the manipulated
of the heating system is approximated by a first order model. variables feed and reflux ratio have minimised condition
C. Noeres et al. / Chemical Engineering and Processing 43 (2004) 421–434 431

Table 7 [29]. As described in [24] the minimisation of the integral


Attainable performance properties squared error of the outputs for step changes of the setpoints
Manipulated variables Structure Output tA O Comments can be formulated as a quadratic program. In this formula-
(h) (%) tion input limitations and steady-state accuracy are included
Heat duty/reflux ratio {3 22} y1 0.17 9 – as linear constraints. In Table 7 the main properties of the
y2 0.02 3 – attainable performance are summarised for the remaining
Heat duty/reflux ratio {3 10} y1 0.23 2 – best structures.
y2 0.08 4 – From an engineering point of view it is not surprising that
Heat duty/reflux ratio {3 11} y1 0.23 3 – the structure {3 22} shows a good performance as the heat
y2 0.05 2 Coupled duty directly determines the vapour flow and therefore the
HAc-feed/reflux ratio {3 13} y1 0.34 4 – molar flow rate of the product. This signifies that one degree
y2 0.16 1 – of freedom is only substituted by another one. From this it
HAc-feed/reflux ratio {3 14} y1 0.27 7 – follows that the resulting control structure may be considered
y2 0.31 1 – as a single input single output structure. The second degree
HAc-feed/reflux ratio {3 4} y1 0.23 2 –
of freedom, e.g. the vapour flow, is not used. Regarding the
y2 0.16 2 – location of the temperature measurements 13 and 14 it is
also obvious that these temperatures can be controlled easily
tA : rise time, O: overshoot.
by the HAc-feed. For this reasons the structures {3 22}, {3
13} and {3 14} in Table 7 are omitted.
numbers γ ∗ > 10. These structures are not considered in For the remaining structures, a linear controller design
the further selection process. is carried out using multivariable frequency response ap-
The criteria discussed above give only qualitative indica- proximation [30]. The nominal desired closed-loop systems
tions whether a control structure is likely to lead to good are defined by the results of the computation of the attain-
controller performance or not. For a quantitative investiga- able performance. The resulting linear MIMO controllers
tion, the promising candidate structures must be compared were applied to the non-linear RBA model. In Fig. 9, the
directly with respect to their attainable performance. A rig- disturbance rejection for the structure {3 11} is depicted
orous comparison is possible under mild assumptions on the for the disturbance scenario summarised in Table 8. It can
problem setting using convex optimisation [27,28]. The key be concluded that the developed controller is robust against
idea is to use the Youla parameterisation of all closed-loop the given uncertainties and shows good performance as pre-
systems which is affine in a free transfer matrix, and to op- dicted by the linear analysis. The rigorous RBA model used
timise the free parameter for convex performance criteria in the simulation differs from the model from which the

Fig. 9. Performance of the controller for the structure {3 11} with the manipulated variables heat duty and modified reflux ratio.
432 C. Noeres et al. / Chemical Engineering and Processing 43 (2004) 421–434

Table 8 For the problem at hand the profit is maximised. The cost
Disturbances to the process function f is defined as:
nMeAc
Disturbance Nominal First Second Time f = . (12)
value value value (h) tf + tstartup + tshutdown
Heat duty (kW) 3.5 0 −1 3:10
Modified reflux ratio 0.6 0 −0.1 4:40 The decision variables are the setpoints of the methyl acetate
Factor reaction rate 1 1 0.5 6:10 fraction in the reflux and the molar product flow, the feed
Heat loss per stage (W) 0 0 10 7:40 rate of acetic acid and the final time of the batch. To ensure
Feed (mmol/s) 39 0 −10 9:10 that the column is operated within its limits upper and lower
bounds to the manipulated variables are introduced. A lower
bound of the mole fraction of methyl acetate in the distillate
linearisation was computed. So the given results also demon- tank is prescribed as well as lower bounds for the conversion
strate the robustness of the controller. of methanol and acetic acid are defined. Thus the closed-loop
optimisation problem is formulated as:
5.2. Optimisation max
s
f (13)
ti ,yi ,Fi
The robust controller is used in the optimisation of the
subject to the plant dynamics
process in a closed-loop fashion. In contrast to the open-loop
optimisation which was investigated by Fernholz et al. [25], F(ẋ(t), x(t), y(t), u(t)) = 0 (14)
in the closed-loop approach the setpoints of the control vari- and the constraints
ables are used as free parameters instead of the manipulated
Qi,min ≤ Qi ≤ Qi,max
variables. By applying the open-loop optimisation process
it is not assured that a linear controller is able to track RTi,min ≤ RTi ≤ RTi,max
the computed trajectory due to changes in the steady-state Fi,min ≤ Fi ≤ Fi,max
gain of the process. If the controller action is considered xMeAc
distillate ≥ xmin
(15)
MeAc
in the optimisation, the controllability of the process with
the controller is maintained. This is the steady-state gain of CMeOH (tf ) ≥ CMeOH
min

the process may change its sign so that linear controllers CHAc (tf ) ≥ CHAc
min

fail. Setpoints which lead to this situation are therefore not N


i=1 ti = tf
optimal and are thus penalised.
In the literature different formulations of the optimisation where yis are the setpoints of the control variables. Due
problem for batch distillation columns can be found [31–33]. to the dynamic equality constraints, a dynamic non-linear

Fig. 10. Comparison between open-loop and closed-loop optimisation.


C. Noeres et al. / Chemical Engineering and Processing 43 (2004) 421–434 433

optimisation problem results. It is solved by transformation Acknowledgements


to a non-linear program using control vector parameterisa-
tion [25]. The time horizon is divided into five intervals in We would like to thank our colleagues of the DFG
which the free variables are parameterised functions of time, research group “Integrated Reaction and Separation
here piecewise constant values are taken. The length of the Processes” for the fruitful co-operation. We are grateful to
intervals are free variables. The optimisation is performed the German Research Foundation (DFG, Grant No. Schm
using the SQP-based algorithm provided by gOPT [22]. 808/5) for financial support. Moreover, we would like to
Fig. 10 shows a comparison between the two optimisa- express our thanks to the “Fond der Chemischen Industrie”,
tion strategies. For the start up interval in both cases the Bayer AG and Celanese AG for providing chemicals.
open-loop approach was used as in that interval changes in
the gain are inevitable and therefore a linear controller must
fail. Appendix A. Notations
From Fig. 10 it follows that the calculated trajectories
of the manipulated variables are similar except for the last a modified activity coefficient
interval. In the closed-loop formulation constant values for acat specific gas-liquid interfacial area (m2 /m3 )
the methyl acetate fraction are stipulated in the different time ai specific gas-liquid interfacial area (m2 /m3 )
intervals. Regarding the trajectory of the heat duty, the drift Ac column cross-section (m2 )
of the process can be recognised. c molar concentration (mol/m3 )
C conversion
Dax,L axial dispersion coefficient (m/s2 )
6. Conclusions and outlook E specific energy holdup (J/m)
EA activation energy (J/mol)
Model-based design, control and optimisation issues of h molar enthalpy (J/mol)
catalytic distillation processes were discussed in this contri- HR0 standard reaction enthalpy (kJ/mol)
bution. (J) diffusion flux vector (mol/(m2 s))
A rigorous rate-based approach has been developed which k reaction rate coefficient (mol/(g s))
describes packing characteristics as liquid holdup, liquid K binary adsorption coefficient
backmixing and pressure drop as well as reaction kinetics. L liquid molar flow rate (mol/s)
Extensive pilot plant experiments for methyl acetate syn- mcat catalyst amount (g)
thesis have been performed and transient concentration and n molar amount (mol)
temperature profiles were used to validate the dynamic pro- nc number of components of mixture
cess model. N molar flux (mol/(m2 s))
The results presented in this contribution demonstrate Q heat flux (W/m2 )
clearly that the suggested rigorous rate-based model reflects ri reaction rate (mol/s)
the dynamic and pseudo steady-state behaviour of the re- [R] inverse mass transfer coefficient matrix (s/m)
garding RD process. The posed model is applicable to pro- RT modified reflux ratio, ν/(ν − 1)
cess design and scale-up issues. Moreover, it was used as a t time (s)
reference for simplifications of the model. T temperature (K)
These reduced models are necessary for control and opti- uL liquid velocity (m/s)
misation purposes as the rigorous model is not suitable for U specific molar holdup (mol/m)
these tasks due to its complexity. Based on a linearisation V vapour molar flow rate (mol/s)
of a simplified equilibrium stage model including a reac- x liquid mole fraction (mol/mol)
tion kinetics term a systematic control structure selection y vapour mole fraction (mol/mol)
procedure was applied. The resulting controller was veri- z axial co-ordinate (m)
fied using the complex rate-based model and shows good
performance and robustness to model uncertainties and dis- Greek letters
turbances. Using the derived linear controller a closed-loop αax thermal dispersion coefficient (m2 /s)
optimisation based on the equilibrium stage model was per- β mass transfer coefficient (m/s)
formed. This approach ensures the controllability of the pro- γ∗ minimised condition number
cess for the calculated trajectories. But as linear controllers [Γ ] thermodynamic correction matrix
are not able to drive the process in arbitrary regions of oper- δ film thickness (m)
ation, non-linear model-based controllers should be applied. λ Molecular thermal conductivity (W/m K)
For the necessary effective simulation of the process local σ̄ mean relative error
linear models interpolated by radial basis functions will be υ stoichiometric coefficient
used. The developed control strategies will be applied to the χ mole fraction
pilot plant in order to verify the developed concepts. ωN band width of rhp zero (rad/s)
434 C. Noeres et al. / Chemical Engineering and Processing 43 (2004) 421–434

Subscripts [16] J. Mackowiak, Fluiddynamik in Kolonnen mit modernen Füllkörper-


packungen und Packungen für Gas/Flüssigkeitssysteme, Sauerländer,
i, j, k component indices
Frankfurt am Main, 1991.
t total [17] V.H. Agreda, L.R. Partin, W.H. Heise, High-purity methyl acetate
via reactive distillation, Chem. Eng. Progress 86 (2) (1990) 40–46.
Superscript [18] W. Song, G. Venimadhavan, J.M. Manning, M.F. Malone, M.F. Do-
B bulk phase herty, Measurement of residue curve maps and heterogeneous ki-
netics in methyl acetate synthesis, Ind. Eng. Chem. Res. 37 (1998)
1917–1928.
[19] T. Pöpken, L. Götze, J. Gmehling, Reaction kinetics and chemical
equilibrium of homogeneously and heterogeneously catalyzed acetic
References acid esterification with methanol and methyl acetate hydrolysis, Ind.
Eng. Chem. Res. 39 (2000) 2601–2611.
[1] M.F. Doherty, G. Buzad, Reactive distillation by design, Trans. [20] J. Gmehling, B. Kolbe, Thermodynamik, Verlag, Weinheim, 1992.
IChemE 70 (1992) 448–458. [21] R. Reid, J.M. Prausnitz, B.E. Poling, The Properties of Gases and
[2] R. Taylor, R. Krishna, Modelling reactive distillation, Chem. Eng. Liquids, McGraw-Hill, New York, 1987.
Sci. 55 (2000) 5183–5229. [22] gPROMS Introductory User Guide, Release 1.84, Process Systems
[3] C. Noeres, E.Y. Kenig, A. Górak, Modelling of reactive separation Enterprise Ltd., UK, 2000.
processes: reactive absorption and reactive distillation, Chem. Eng. [23] P. Moritz, H. Hasse, Fluid dynamics in reactive distillation packing
Process. (2003) 159–178. katapak-S, Chem. Eng. Sci. 54 (1999) 1367–1374.
[4] A. Górak, L.U. Kreul, M. Skowronski, Strukturierte Mehrzweck- [24] S. Pegel, S. Engell, Control Structure Selection for an Air Separation
packung, German Patent 19701045 A1. Plant—A Systematic Approach, IFAC Symposium DYCOPS 6, Cheju
[5] Sulzer, KATAPAK (product information), Sulzer Chemtech AG, Island, Korea, 2001.
2000. [25] G. Fernholz, S. Engell, L.-U. Kreul, A. Górak, Optimal operation of
[6] A. Górak, A. Hoffmann, Catalytic distillation in structured packings: a semi-batch reactive distillation column, Comput. Chem. Eng. 24
methyl acetate synthesis, AIChE J. 47 (2001) 1067–1076. (2000) 1569–1575.
[7] C. Thiel, Modellbildung, Simulation, Design und experimentelle Va- [26] S. Engell, Controllability analysis and control structure selection,
lidierung von heterogen katalysierten Reaktivdestillationsprozessen Invited paper, IFAC Symposium on New Trends in Design of Control
zur Synthese der Kraftstoffether MTBE, ETBE und TAME, Ph.D. Systems, 1997, pp. 61–68.
Thesis, Technical University of Clausthal, Germany, 1997. [27] S. Boyd, C. Barrat, Linear Controller Design: Limits and Perfor-
[8] S. Engell, G. Fernholz, Control of a reactive separation process, mance, Prentice Hall, Englewood Cliffs, 1991.
Chem. Eng. Process. 42 (2002) 201–210. [28] K. Webers, Controller design via an lmi-based stability test and
[9] G. Fernholz, Prozeßführung einer halbkontinuierlich betriebenen convex optimization, in: Proceedings of the 2nd IFAC Symposium on
Kolonne zur Reaktivrektifikation, Ph.D. Thesis, University of Dort- Robust Control Design, ROCOND ‘97, Budapest, 1997, pp. 223–228.
mund, Germany, 2000. [29] K. Webers, Bestimmung der erreichbaren Regelgüte durch konvexe
[10] L.-U. Kreul, Discontinuous and Reactive Distillation, Ph.D. Thesis, Optimierung, Schriftenreihe des Lehrstuhls für Anlagensteuerung-
University of Dortmund, Germany, 1998. stechnik, Bd. 3/97, Shaker Verlag, Aachen, Dissertation, Universität
[11] R. Taylor, R. Krishna, Multicomponent Mass Transfer, Wiley, New Dortmund, 1997.
York, 1993. [30] R. Müller, FASTER, Frequenzgangapproximation zur Synthese von
[12] A. Górak, Simulation thermischer Trennverfahren fluider Vielkom- Regelkreisen für MATLAB, Technischer Bericht Version 2.01, Uni-
ponenten-gemische, in: G. Schuler (Ed.), Prozeßsimulation, VCH versität Dortmund, 1995, Copyright 1990–1995.
Verlag, Weinheim, 1995, pp. 349–408. [31] U.M. Diwekar, R.K. Malik, K.P. Madhavan, Optimal reflux rate
[13] N. Wakao, S. Kaguei, Heat and Mass Transfer in Packed Beds, policy determination for multicomponent batch distillation columns,
Science Publishers, Gordon and Breach, 1992. Chem. Eng. Progress 88 (3) (1987) 43–50.
[14] S. Pelkonen, Multicomponent Mass Transfer in Packed Distilla- [32] H. Egly, V. Ruby, B. Seid, Optimum design and operation of batch
tion Columns, Ph.D. Thesis, University of Dortmund, Germany, rectification accompanied by chemical reaction, Comput. Chem. Eng.
1997. 3 (1979) 169–174.
[15] C. Noeres, A. Hoffmann, A. Gorak, Reactive distillation: non-ideal [33] J.S. Logsdon, U.M. Diwekar, L.T. Biegler, On the simultaneous
flow behaviour of the liquid phase in structured catalytic packings, optimal design and operation of batch distillation columns, Chem.
Chem. Eng. Sci. 57 (2002) 1545–1549. Eng. Res. Des. 68 (1990) 434–444.

You might also like