Top Banner

of 21

1-s2.0-S0009250904004282-main

Apr 14, 2018

Download

Documents

Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
  • 7/27/2019 1-s2.0-S0009250904004282-main

    1/21

  • 7/27/2019 1-s2.0-S0009250904004282-main

    2/21

    4548 I.L. Chien et al. / Chemical Engineering Science 59 (2004) 45474567

    x (H O)

    0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

    y(HO

    )

    0.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    x (H O)

    0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

    RelativeVolatility((yHO

    /xHO

    )/(y

    /x

    ))

    1.0

    1.2

    1.4

    1.6

    1.8

    2.0

    2.2

    Fig. 1. VLE and relative volatility for HAc and water system.

    alcohol + cyclohexane + water system. Experimental veri-

    fication of the above control strategy can be seen in Chien

    et al. (2000a). Chien et al. (2000b) proposed a simple oper-

    ating procedure under inverse double loop control strategy

    to automatically adjust the column heat duty and organic re-

    flux to be at optimum operating point and also proposed an

    improved decoupling control strategy for the double loops.

    Ulrich and Morari (2002) examine the influence of fourth

    component impurities on the operation and control of a het-

    erogeneous azeotropic distillation column for dewatering a

    heavy-boiling organic using methyl tert-butyl ether as a lightentrainer. None of the above papers studied the acetic acid

    dehydration system.

    Design of acetic acid dehydration system using an en-

    trainer has been studied in several publications. In a review

    paper, Othmer (1963) described an azeotropic distillation

    system containing a dehydrating column, a decanter, and a

    water column for the separation of acetic acid and water.

    The entrainer used before 1932 was ethylene dichloride, and

    later normal propyl acetate and normal butyl acetate were

    used to reduce the organic reflux and heat duty used in the

    dehydrating column. In the paper by Pham and Doherty

    (1990), examples of using ethyl acetate (cf. Tanaka and

    Yamada, 1972), n-propyl acetate (cf. Othmer, 1941), or n-

    butyl acetate (cf. Othmer, 1941; Tanaka and Yamada, 1972)

    as entrainer were listed in a table of examples of heteroge-

    neous azeotropic separations. Siirola (1995) uses acetic acid

    dehydration as an example to demonstrate a systematic pro-

    cess synthesis technique to the conceptual design of process

    flowsheet. Ethyl acetate as entrainer was used in the paperby Siirola (1995) to design a complete acetic acid dehydra-

    tion process with multiple effect azeotropic distillation and

    heat integration. More recently, Wasylkiewicz et al. (2000)

    proposed using geometric method for optimum process de-

    sign of an acetic acid dehydration column with n-butyl ac-

    etate as entrainer.

    All of the above papers on acetic acid dehydration system

    are on the subject of process synthesis and design, very lit-

    tle discussion about control strategy of this system has been

    found in the literature. Luyben and Tyreus (1998) offered

    a realistic vinyl acetate monomer example for academic re-

    searchers pursuing simulation, design, and control studies.

    In this example, an azeotropic distillation column with de-

    canter is presented. Although the flowsheet of this column

    system is similar to this study with components of acetic acid

    and water, but since vinyl acetate is a product of the overall

    process, an extra organic phase product is drawn-off from

    the decanter which is different from the system which will be

    studied in this paper. Kurooka et al. (2000) proposed a non-

    linear control system for the acetic acid dehydration column

    with n-butyl acetate as entrainer. The thermodynamic model

    used in this work is questionable because a minimum-boiling

    azeotrope is predicted between n-butyl acetate and acetic

    acid though the mixture is zeotropic (cf. Horsley, 1973).

    In their study, complicated exact inputoutput linearizationcontroller was used with values of some unmeasured state

    variables needed for the calculations. The resulting control

    performances under feed rate and composition changes are

    not desirable because of large fluctuations in the manipu-

    lated variables. Gaubert et al. (2001) studied operation of an

    unnamed organic acid dehydration in the industry using an

    immiscible entrainer. Multiple steady states are confirmed

    for the heterogeneous column by simulation and experimen-

    tal data for the industrial unit. However, dynamics and con-

    trol of this system is not studied in their paper.

    In this paper, a suitable entrainer for this acetic acid de-

    hydration system will be selected from several candidate ac-etates. Steady-state tray by tray column simulation will be

    used to determine the best entrainer with minimum total an-

    nual cost. Optimum process design and operating condition

    will be determined to keep high-purity bottom acetic acid

    composition and also keep a small acetic acid loss through

    top aqueous draw. The overall control strategy of this col-

    umn system will be proposed to hold both bottom and top

    product specifications in spite of feed rate and feed compo-

    sition disturbances. In the control study, conventional con-

    trol strategy using only tray temperature measurements will

    be considered so that the result of this study can easily be

    used directly in industry.

    http://-/?-http://-/?-http://-/?-
  • 7/27/2019 1-s2.0-S0009250904004282-main

    3/21

    I.L. Chien et al. / Chemical Engineering Science 59 (2004) 4547 4567 4549

    Table 1

    Experimental physical properties of three candidate acetates

    Acetate Normal Azeotropic Azeotropic Azeotropic Aqueous Organic

    boiling component temp. composition phase phase

    point ( C) ( C) (water, mol%) (acetate, mol%) (acetate, mol%)

    Ethyl 77.15 Water 70.38 31.2% 1.40% 83.5%

    acetate (@40 C) (@40 C)iso-butyl 117.2 Water 87.4 56.1% 0.127% 90.47%

    acetate (@40 C) (@40 C)

    n-butyl 126.2 Water 90.2 72.2% 0.3638% 86.31%

    acetate (@25 C) (@25 C)

    Table 2

    Parameter values for the NRTL model

    i, j Bij Bj i ij

    1, 2 576.234 322.424 0.3

    1, 3 416.124 1024.50 0.3067

    2, 3 211.310 652.995 0.3

    (1) Ethyl acetate, (2) acetic acid, (3) water.

    2. Process simulation and entrainer selection

    Three candidate acetates will be studied in detailed pro-

    cess simulation to demonstrate the factors needed to be

    considered in determining the suitable entrainer for this

    system. The three candidate acetates to be considered are:

    ethyl acetate, iso-butyl acetate, and n-butyl acetate. The

    important experimental physical properties of these three

    acetates at atmospheric pressure are listed in Table 1. The

    azeotropic data is from Horsley (1973), the vaporliquidequilibrium data is from Gmehling and Onken (1977) with

    the VLE date for acetic acidiso-butyl acetate system from

    Christensen and Olson (1992). For the binary and ternary

    liquidliquid equilibrium data, they are from SZrensen and

    Arlt (1979, 1980). The nonrandom two-liquid (NRTL) ac-

    tivity coefficient model (Renon and Prausnitz, 1968) was

    used for the vaporliquidliquid equilibrium (VLLE) for

    the ternary system. The HaydenOConnell (Hayden and

    OConnell, 1975) second virial coefficient model with asso-

    ciation parameters was used to account for the dimerization

    of acetic acid in the vapor phase. The Aspen Plus (Aspen

    Technology, Inc., 2001) built-in association parameters wereemployed to compute fugacity coefficients. The extended

    Antoine equation is used to calculate the vapor pressure of

    each component in the system. The Aspen Plus built-in

    parameters were again used in the simulation. The set of

    NRTL parameters are obtained to be capable of describing

    well the binary and ternary, vaporliquid equilibrium (VLE)

    and liquidliquid equilibrium (LLE) data. The set of NRTL

    parameters for the ternary systems of acetic acidethyl

    acetatewater, acetic acidiso-butyl acetatewater, and

    acetic acidn-butyl acetatewater are listed in Tables 24.

    All three candidate entrainers form a minimum-boiling

    azeotrope with water. A heterogeneous azeotropic distilla-

    Table 3

    Parameter values for the NRTL model

    i, j Bij Bj i ij

    1, 2 194.416 90.268 0.3

    1, 3 489.609 1809.079 0.2505

    2, 3 211.310 652.995 0.3

    (1) Iso-butyl acetate, (2) acetic acid, (3) water.

    Table 4

    Parameter values for the NRTL model

    i, j Bij Bj i ij

    1, 2 397.85 68.61 0.3

    1, 3 354.31 2578.35 0.219

    2, 3 211.31 652.995 0.3(1) n-Butyl acetate, (2) acetic acid, (3) water.

    tion column can be designed to obtain high-purity acetic acidproduct (b.p. of 118 C) at the column bottom while obtain-

    ing minimum boiling entrainerwater azeotrope at the top

    of the column. With this column design by adding entrainer

    into the system, the difficult tangent pinch of the pure wa-

    ter side can be avoided at the top of the column. Since this

    entrainerwater azeotrope is heterogeneous, the top column

    vapor stream forms two liquid phases after condensation in

    the decanter. The organic phase will be refluxed back to

    the heterogeneous azeotropic column to provide enough en-

    trainer inside of the column. The aqueous phase containing

    mostly water will be assumed to be drawn out from the sys-

    tem for further treatment or discharge. Some of the aqueousphase can be refluxed back to the heterogeneous azeotropic

    column if the organic reflux is too small to fulfill the column

    specifications. The conceptual design of this heterogeneous

    azeotropic distillation column system is illustrated in Fig. 2.

    The residue curve maps with the binodal curve of the

    LLE of the three entrainer systems studied in this paper

    are shown in Figs. 35. By observing these three figures,

    the prediction for entrainer solubility in water and also the

    azeotropic temperature match well with the experimental

    data in Table 1. The azeotropic composition for the iso-butyl

    acetate system gives the most discrepancy in comparison

    with the experimental data in Table 1. The experimental

    http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-
  • 7/27/2019 1-s2.0-S0009250904004282-main

    4/21

    4550 I.L. Chien et al. / Chemical Engineering Science 59 (2004) 45474567

    Fig. 2. Conceptual design for the separation of acetic acid and water.

    Fig. 3. Residual curve map for the system HAcEtAcH2O.

    azeotropic composition is at 43.9mol% iso-butyl acetate but

    the simulation predicted at 36.8mol%. This is mainly due to

    the compromise of obtaining the NRTL model parameters by

    fitting all binary and ternary VLE and LLE data while trying

    to predict well the azeotropic temperature and composition.

    The residue curve maps for the ethyl acetate and the iso-

    butyl acetate systems are similar in nature with the two-

    componentazeotropeas the lowest temperature in thesystem

    and acetic acid as the highest temperature in the system. The

    Fig. 4. Residual curve map for the system HAciBuAcH2O.

    residue curve map for the n-butyl acetate system is different

    than the other two systems. In the n-butyl acetate system,

    the highest temperature in the system is n-butyl acetate (b.p.

    126.2 C), not acetic acid (b.p. 118 C). Slippage of entrainer

    into the bottom product stream is the situation needed to

    be avoided for this system. The other two systems do not

    need to worry about this situation because acetic acid is the

    highest temperature in the system which should come out

    of the column through bottom stream in ideal situation.

  • 7/27/2019 1-s2.0-S0009250904004282-main

    5/21

    I.L. Chien et al. / Chemical Engineering Science 59 (2004) 4547 4567 4551

    Fig. 5. Residual curve map for the system HAcnBuAcH2O.

    Rigorous process simulation is performed to find the op-

    timum design and operating conditions of these three en-

    trainer systems. The feed composition of 50 mol% acetic

    acid and 50 mol% water is considered for the Aspen Plus

    simulation. The feed rate is assumed to be 500kg/h and it

    is saturated liquid phase. The column pressure is assumed

    to be at atmospheric pressure. The decanter temperature is

    at 40 C. In the Aspen Plus simulation, the column bot-

    tom product is kept at 99.9mol% acetic acid high purity by

    varying the reboiler heat duty and the column top aqueousproduct is kept at 0.1 mol% acetic acid loss by varying the

    entrainer makeup flow rate. If the high purity specifications

    cannot be met, portions of the aqueous phase can be refluxed

    back to the column to fulfill the column specifications. This

    extra third degree of freedom (aqueous reflux flow rate) is

    fixed at a value which will meet both top and bottom prod-

    uct specifications while also minimize reboiler heat duty of

    the column system.

    Design variable of total number of trays is a compromise

    between the total equipment cost and the total utility cost.

    The optimum total number of trays and the feed tray location

    are determined to minimize Total Annual Cost (TAC). Thecalculation procedure of Douglas (1988) is followed with

    the annual capital charge factor of 1/3 was used. The utility

    cost is calculated the same way as in Chiang et al. (2002).

    The Aspen Plussimulation results for the three entrainers

    are summarized in Tables 57.

    Several observations can be made by comparing these

    three tables. Firstly, for the system of acetic acidethyl

    acetatewater, no aqueous reflux is necessary to meet the

    product specifications while the other two systems need

    aqueous reflux stream for the separation with higher aque-

    ous reflux flow rate for the n-butyl acetate system. Secondly,

    the organic reflux flow rate and also the reboiler heat duty

    for the ethyl acetate system are much larger in compari-

    son with the other two systems. This high organic reflux

    flow rate in the ethyl acetate system can actually be pre-

    dicted by the inner molar balance envelope in Fig. 2 with

    the residue curve map plot of the ethyl acetate system in

    Fig. 3. Assuming at ideal condition, the column top vapor

    composition should be at the ethyl acetatewater azeotropeand the column bottom composition should be very close

    to the pure acetic acid corner in Fig. 3. Because of the feed

    composition is at 50 mol% acetic acid and 50 mol% water

    and the other inlet stream to the column for the inner molar

    balance envelope in Fig. 2 is the organic reflux (recall that

    no aqueous reflux is necessary for this system), the inter-

    ception of the two inlet and outlet molar balance lines can

    be used to estimate the organic reflux flow rate. Since the

    interception point is closer to the organic reflux composi-

    tion point, the organic reflux flow rate is quite high. If the

    feed is much richer in acetic acid, the organic reflux flow

    rate will be lower than the current case.

    Another observation by comparing these three tables is

    that the makeup flow rate for the ethyl acetate system is the

    highest while for the iso-butyl acetate system is the lowest.

    This can be explained by the outer molar balance envelope

    in Fig. 2 with the knowledge of the aqueous phase composi-

    tion in Fig. 3. Assuming ideal situation for the ethyl acetate

    system, the two outlet streams for the outer molar balance

    envelope are at the points of aqueous phase composition and

    pure acetic acid in Fig. 3. The two inlet streams are at the

    points of feed composition and pure ethyl acetate (entrainer

    makeup) point. How close the interception point of the two

    molar balance lines to the feed composition point can be

    used to determine the makeup flow rate since the feed flowrate is known. If this interception point is very close to the

    feed composition point, the makeup flow rate will be small.

    From this explanation, it is not difficult to conclude that the

    ethyl acetate system will have the highest makeup flow rate

    and the iso-butyl acetate system will have the lowest makeup

    flow rate.

    The comparison of the minimum attainable TAC for these

    three systems as well as the acetic acid dehydration system

    without any entrainer is shown in Table 8. From the table,

    one can observe that the no entrainer system required the

    most TAC and the iso-butyl acetate system is most favor-

    able for this feed composition and product specification re-quirements. The TAC for the iso-butyl acetate system is only

    about 55% of the no entrainer system which represents large

    saving can be made by using the iso-butyl acetate system.

    Notice that this finding is in general agreement with the in-

    dustrial applications. (See patents by Costantini et al., 1981

    and by Parten and Ure, 1999). The above two patents also

    found iso-butyl acetate as a favorable entrainer for the sep-

    aration of acetic acid and water. Another earlier patent by

    Othmer (1936) found n-propyl acetate to be useful as an en-

    trainer for this system. The patent by Mitsui Petro-Chemical

    Industries, Ltd. (1980) found n-butyl acetate to be favor-

    able for this system. Notice that in all the patents above, the

    http://-/?-http://-/?-http://-/?-
  • 7/27/2019 1-s2.0-S0009250904004282-main

    6/21

    4552 I.L. Chien et al. / Chemical Engineering Science 59 (2004) 45474567

    Table 5

    Stream information for the system acetic acid(HAc)ethyl acetate(EtAc)water(H2O)

    Feed Bottom Top Aqueous Organic Makeup Reboiler

    product product reflux reflux stream energy

    Flow rate 213.48 106.74 108.98 0 573.95 2.24

    (mol/min)

    HAc mole 0.5 0.999 1.00103 2.87103 0 fraction

    H2O mole 0.5 9.90104 0.9785 0.17586 0

    fraction

    EtAc mole 0.0 1.0105 2.05102 0.82127 1

    fraction

    Heat duty (KW) 401.17

    Table 6

    Stream information for the system acetic acid(HAc)iso-butyl acetate(iBuAc)water(H2O)

    Feed Bottom Top Aqueous Organic Makeup Reboiler

    product product reflux reflux stream energy

    Flow rate 213.48 106.74 106.90 33.36 92.71 0.16

    (mol/min)

    HAc mole 0.5 0.999 1.00103 1.00103 1.90103 0

    fraction

    H2O mole 0.5 5.90104 0.9979 0.9979 7.98102 0

    fraction

    iBuAc mole 0.0 4.10104 1.10103 1.10103 0.9183 1

    fraction

    Heat duty (KW) 167.01

    Table 7

    Stream information for the system acetic acid(HAc)n-butyl acetate(nBuAc)water(H2O)

    Feed Bottom Top Aqueous Organic Makeup Reboilerproduct product reflux reflux stream energy

    Flow rate 213.48 106.74 107.44 98.78 102.32 0.70

    (mol/min)

    HAc mole 0.5 0.999 1.00103 1.00103 1.50103 0

    fraction

    H2O mole 0.5 7.48104 0.9928 0.9928 0.1448 0

    fraction

    nBuAc mole 0.0 2.52104 6.20103 6.20103 0.8537 1

    fraction

    Heat duty (KW) 259.68

    Table 8Comparison of total annual cost for the acetic acid dehydration systems

    Entrainer Optimal Optimal Annualized Utility cost Entrainer TAC($)

    total feed capital cost cost

    stages stage

    Ethyl acetate 16 2 6.84104 4.20104 5.40104 1.64105

    iso-butyl 30 9 6.81104 1.80104 1.70104 1.03105

    acetate

    n-butyl 31 11 8.44104 2.78104 6.08104 1.73105

    acetate

    No entrainer 50 37 1.42105 4.37104 0 1.86105

  • 7/27/2019 1-s2.0-S0009250904004282-main

    7/21

    I.L. Chien et al. / Chemical Engineering Science 59 (2004) 4547 4567 4553

    Fig. 6. Vapor and liquid profiles for the optimum system HAc

    iBuAcH2O.

    designed feed compositions and the specified column bot-

    tom and top purities are all different from this paper, thus

    no direct comparison of results can be made.

    The vapor and liquid profiles inside the column for the

    optimized iso-butyl acetate system can be seen in Fig. 6.

    Notice that the column behaves as what was designed. The

    first five stages counting from the top of the column have two

    liquid phases. The combined liquid compositions for thesefive stages are plotted in this figure. Another thing worth

    mention in the figure is that the column composition profiles

    bypassing the corner of pure water which is the region for

    the tangent pinch to be avoided.

    From this study, some important factors in determining the

    suitable entrainer for the acetic acid dehydration system are

    summarized below. The information needed for this qualita-

    tive comparison can be illustrated by the residue curve maps

    with the binodal curve of the LLE as shown in Figs. 35.

    The suitability of the entrainer is actually a combination of

    the following factors.

    2.1. Azeotropic composition and organic phase composition

    It is better to have the azeotropic composition containing

    more water in this mixture. This means that this entrainer is

    more capable of carrying water to the top of the column, thus

    less entrainer is needed inside of the column. The distance

    for the points between azeotropic composition and organic

    phase composition is better to be further apart. This means

    that besides that the azeotropic composition containing more

    water, the organic phase composition should contain more

    entrainer. The location of these two points in Figs. 35 have

    to do with the organic reflux flow rate into the heterogeneous

    column as explained above during estimating the organic

    reflux flow rate for the ethyl acetate system. From Figs. 35,

    the ethyl acetate is the worst entrainer if only considering

    this factor.

    2.2. Azeotropic temperature

    The azeotropic temperature determines the temperature

    difference between the top and the bottom of the column. A

    large delta T of the azeotropic temperature to the pure acetic

    acid temperature implies a good separability. Less column

    stages will be needed for specific product purity specifica-

    tions. In this regard, ethyl acetate is the best entrainer. This

    interpretation is confirmed by Table 8 because ethyl acetate

    system requires the least total number of stages for the same

    separation.

    2.3. Aqueous phase composition and entrainer pricing

    The aqueous phase should contain as little entrainer as

    feasible. The reason is because the aqueous phase stream

    will be drawn out of the system, thus any entrainer loss

    should be compensated by the makeup stream in Fig. 2.

    This will correspond to a stream cost of the system as seen

    in Table 8. The makeup flow rate can actually be estimated

    using the outer molar balance envelope in Fig. 2 during ideal

    situation as explained previously. In this regard, iso-butyl

    acetate system results in the least makeup flow rate while

    ethyl acetate systemrequires themost makeupflow rate. This

    is confirmed by Tables 57. The annual cost of this stream

    is not only related to its flow rate but also related to theentrainer pricing. In this regard, ethyl acetate is the cheapest

    and iso-butyl acetate is the most expensive entrainer. With

    the knowledge of the entrainer pricing and the calculation

    of the makeup flow rates for the three systems, the entrainer

    cost can be estimated as seen in Table 8 even without any

    rigorous simulation.

    Since the system with iso-butyl acetate as entrainer re-

    sults in most economical process design, we will study the

    dynamic and control strategy of this system in detail in the

    next section. Before doing that, let us first explore the neces-

    sity of the aqueous reflux stream under various feed com-

    positions for the system using iso-butyl acetate as entrainer.Fig. 7 shows the collection of many simulation results under

    various feed composition conditions. In all the simulation

    runs, the total numbers of stages for the column are all fixed

    the same as the one in Table 8 (30 stages including reboiler

    but not the condenser). The column bottom product is kept

    at 99.9 mol% acetic acid high purity by varying the reboiler

    heat duty and the column top aqueous product is kept at

    0.1 mol% acetic acid loss by varying the entrainer makeup

    flow rate. The aqueous reflux flow rate is fixed at value

    which will meet both top and bottom product specifications

    while also minimize reboiler heat duty of this column sys-

    tem. From the figure, one can observed that for feed water

  • 7/27/2019 1-s2.0-S0009250904004282-main

    8/21

    4554 I.L. Chien et al. / Chemical Engineering Science 59 (2004) 45474567

    Feed Water Composition

    0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

    AqueousRe

    fluxFraction

    0.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    Fig. 7. Minimum aqueous reflux ratio under various feed compositions.

    composition above 70 mol%, the aqueous reflux stream is

    not needed. For the feed composition studied in this paper,

    the aqueous reflux stream is necessary in order to properly

    hold the bottom and top product specifications.

    Table 9 shows the value of main operating variables in

    keeping the top and bottom product purity at their specifi-

    cations under various feed composition conditions. In this

    table, the operating condition is not unique for feed water

    contents from 10% to 60%. For these feed composition

    cases, there are three degrees of freedom (extra one is the

    aqueous reflux) for the system with only two product purity

    specifications. The ones included in the table are the operat-

    ing conditions that minimize reboiler heat duty by varying

    aqueous reflux flow rate. One important thing which needs tobe pointed out from the table is that in order to hold product

    specifications, aqueous reflux flow rate needs to be adjusted

    in a very wide range. This means that this manipulated

    variable (aqueous reflux) should not be fixed in the control

    strategy when trying to reject unmeasured feed composition

    disturbance. This manipulated variable is preferable to be

    used in the inferred composition loop to hold product spec-

    ifications. Another observation is that the reboiler duty goes

    Table 9

    Desired operating conditions under various feed compositions

    Feed H2O Aqueous Organic Reboiler Entrainer Aqueous Bottom

    composition reflux reflux duty makeup draw product

    (mol%) (mol/min) (mol/min) (KW) (mol/min) (mol/min) (mol/min)

    10 133 102 189 0.07 21 192

    20 109 100 186 0.08 43 171

    30 85 99 182 0.11 64 150

    40 60 96 175 0.13 86 128

    50 33 93 167 0.16 107 107

    60 8 90 159 0.16 128 85

    70 0 99 175 0.16 150 64

    80 0 113 200 0.19 171 43

    90 0 127 226 0.21 193 21Product specifications: Bottom at 99.9mol% HAc and aqueous draw at 0.1mol% HAc.

    up dramatically for the cases with no aqueous reflux (70%,

    80%, and 90% feed H2O compositions). This implies that

    it may be better to add a preconcentrator column before the

    heterogeneous azeotropic column to increase the acetic acid

    content in the feed to the heterogeneous azeotropic column

    if the fresh feed water composition is too high. Another

    possible advantage may be to have extra degree of freedom(aqueous reflux) to be manipulated in the control strategy.

    The importance of the aqueous reflux stream for manipula-

    tion purpose will be shown in the next section.

    3. Control strategy design

    The heterogeneous azeotropic column system using iso-

    butyl acetate as entrainer will be studied in detail in this

    section. The overall control strategy of this system will be

    developed in order to hold bottom and top product spec-

    ifications in spite of feed flow rate and feed composition

    changes. In the control strategy development, we will as-sume no online composition measurement is available. The

    composition control loops will be inferred by some tray tem-

    perature control strategy. This type of control strategy can

    easily be implemented in industry for wider applications.

    The Aspen Plus steady state simulation in the last

    section is exported to the dynamic simulation of Aspen

    DynamicsTM. The tray sizing option in Aspen Plus is uti-

    lized to calculate the column diameter to be 0.3259m with

    the tray spacing of 0.6096 m is assumed. Other equipment

    sizing recommended by Luyben (2002) is used here. The

    volume of the reboiler is sized to give 10 min holdup with

    50% liquid level. The decanter is sized to be bigger to al-low for two liquid phases to separate. The holdup time of

    20 min is used in the dynamic simulation. Pressure-driven

    simulation in Aspen DynamicsTM is used with the top

    pressure of the azeotropic column controlled at 1.1 atm to

    allow for some pressure drop in the condenser and decanter

    to give the decanter at atmospheric pressure. The pressure

    drop inside the column is automatically calculated in Aspen

    DynamicsTM. Since the tray pressures in the columns are

  • 7/27/2019 1-s2.0-S0009250904004282-main

    9/21

    I.L. Chien et al. / Chemical Engineering Science 59 (2004) 4547 4567 4555

    Table 10

    Base case condition of the optimum flow sheet for dynamic tests

    Total number of stage for the 30 (including reboiler but not

    azeotropic column including condenser)

    Feed stage 9 (counting from the top tray)

    Fresh feed flow rate 213.48 mol/min

    Fresh feed composition 50mol% HAc and 50mol% H2O

    Column reboiler duty 162.27 KW

    Organic reflux flow rate 92.21mol/min

    Aqueous reflux flow rate 33.40mol/min

    Entrainer makeup flow rate 0.165mol/min

    Bottom flow rate 106.76 mol/min

    Bottom composition 99.89 mol% HAc

    0.0665mol% H2O

    0.0458mol% iBuAc

    Top aqueous outlet flow rate 106.89mol/min

    Top aqueous outlet composition 0.0997mol% HAc

    99.79 mol% H2O

    0.11mol% iBuAc

    different than the constant atmospheric pressure assump-

    tion used in steady-state simulation, the established base

    case condition in Aspen DynamicsTM will be slightly dif-

    ferent than Table 6 in previous section. The final base case

    steady-state condition used for control study can be seen in

    Table 10.

    There are two inventory control strategies which can be

    used for this system. The first inventory control strategy (In-

    ventory Strategy #1) uses entrainer makeup flow to control

    theorganic phase level in thedecanter. This inventory control

    strategy was successfully used in Chien et al. (1999b, 2000a)when controlling an isopropyl alcohol dehydration column.

    The second inventory control strategy (Inventory Strategy

    #2) uses organic reflux flow to control the organic phase

    level in the decanter. This second inventory control strategy

    is more intuitively sound because organic reflux flow rate is

    much larger than the entrainer makeup flow rate, thus the

    organic phase level control should be more effective. Other

    inventory control loops which use the same pairings for ei-

    ther of the above strategies are: using top aqueous product

    flow to control the aqueous phase level in the decanter; us-

    ing bottom product flow to control the column bottom level.

    The column top pressure is controlled at 1.1 atm by manip-ulating the top vapor flow and the decanter temperature is

    controlled at 40 C by manipulating the condenser duty.

    After deciding the inventory control strategy, there are

    three variables left and can be used in some composition

    control strategy. The three candidate variables for Inventory

    Strategy #1 are: organic reflux flow, aqueous reflux flow,

    and the reboiler duty; while the three candidate variables for

    Inventory Strategy #2 are: entrainer makeup, aqueous reflux

    flow, and the reboiler duty. The control objective is to hold

    the bottom and the top aqueous product specifications at

    base case condition under 10% feed flow and 10% feed

    H2O composition changes.

    Stages

    1 2 3 4 5 6 7 8 9 101112131415161718192021222324252627282930

    80

    90

    100

    110

    120

    130

    140

    Stages

    1 2 3 4 5 6 7 8 9 101112131415161718192021222324252627282930

    80

    90

    100

    110

    120

    130

    140Stages

    1 2 3 4 5 6 7 8 9 101112131415161718192021222324252627282930

    Te

    mp(oC)

    Temp(oC)

    Temp(

    oC)

    80

    90

    100

    110

    120

    130

    140

    Aqueous Reflux 0.1% Change

    Heat Duty 0.1% Change

    Organic Reflux 0.1% Change

    Base case

    +0.1%

    -0.1%

    Base case

    +0.1%

    -0.1%

    Base case

    +0.1%

    -0.1%

    Fig. 8. Sensitivity analysis of the three manipulated variables under Strat-

    egy #1.

    3.1. Dual temperature loop control strategy

    Since product specifications at both bottom and top ends

    are specified, we will consider dual-point temperature con-

    trol structure first. The sensitivity analysis with small pertur-

    bation of the manipulated variables will be performed next

    in order to determine the two temperature control points.

    Fig. 8 shows the sensitivity analysis of the three manipulated

    variable changes under Inventory Strategy #1 and Fig. 9

    shows the sensitivity analysis of the three manipulated vari-

    able changes under Inventory Strategy #2. The numbering

    of the stage in this column is counting from top to bottom

    http://-/?-http://-/?-
  • 7/27/2019 1-s2.0-S0009250904004282-main

    10/21

    4556 I.L. Chien et al. / Chemical Engineering Science 59 (2004) 45474567

    Stages

    1 2 3 4 5 6 7 8 9 1 0 1 1 12 1 3 1 4 1 5 16 1 7 1 8 1 9 20 2 1 2 2 2 3 24 2 5 2 6 2 7 28 2 9 3 0

    T

    emp(C)

    80

    90

    10 0

    11 0

    12 0

    13 0

    14 0

    Base Case

    +0.1%

    -0.1%

    Aqueous Reflux 0.1% Change

    Stages

    1 2 3 4 5 6 7 8 9 10 1 1 1 2 13 1 4 1 5 16 1 7 1 8 19 2 0 2 1 22 2 3 2 4 25 2 6 2 7 28 2 9 3 0

    Temp(C)

    80

    90

    10 0

    11 0

    12 0

    13 0

    14 0

    Heat Duty 0.1% Change

    Stages

    1 2 3 4 5 6 7 8 9 10 1 1 1 2 13 1 4 1 5 16 1 7 1 8 19 2 0 2 1 22 2 3 2 4 25 2 6 2 7 28 2 9 3 0

    Temp

    (C)

    80

    90

    10 0

    11 0

    12 0

    13 0

    14 0

    Base Case

    +10%

    -10%

    Makeup Flow 10% Change

    Base Case

    +0.1%

    -0.1%

    Fig. 9. Sensitivity analysis of the three manipulated variables under Strat-

    egy #2.

    with stage #1 as the top stage and stage #30 as the reboiler.When perturbing one manipulated variable, the other two

    manipulated variables are fixed at base case condition. The

    final steady-state conditions of Figs. 8 and 9 are obtained

    by running dynamic simulation with the above mentioned

    perturbations and then wait until the dynamic simulation to

    reach final steady-state.

    For organic reflux changes in Fig. 8, a process gain sign

    reversing is observed between stages #12 and #13. Dynam-

    ically, a large inverse response was observed for column

    stages between stages #13 to column bottom. Similarly for

    entrainer makeup changes in Fig. 9, a process gain sign re-

    versing is also observed between stages #12 and #13. This

    Stages

    5 10 15 20 25 30

    Zi=|U1i|-|U2i|

    -0.2

    -0.1

    0.0

    0.1

    0.2

    0.3

    Fig. 10. Plot ofZi for control structure CS1.

    also indicates that dynamically a large inverse response was

    also observed for column stages between stages #13 to col-

    umn bottom.

    Since the numbers of the candidate manipulated variables

    for each inventory control strategy are three, there will be

    three alternative temperature control structures for each in-

    ventory control strategy. From the open-loop data in Figs. 8

    and 9, steady-state gain matrix for each alternative overall

    control strategy can be obtained by averaging the positive

    and negative manipulated variable changes. Each elements

    of the steady-state gain matrix is made to be dimensionless

    by the spans of the temperature sensors and the manipulated

    variables. Singular-value decomposition (SVD) as described

    by Moore (1992) can be made on the steady-state gain ma-trices as follows:

    K= UVT, (1)

    where K is a 302 steady-state gain matrix for each control

    strategy. U= [U1|U2] is an 302 orthonormal matrix, the

    columns of which are called the left singular vectors.VT is a

    22 orthonormal matrix, the columns of which are called the

    right singular vectors. is a 22 diagonal matrix of scalars

    called the singular values and organized in descending order.

    To trade off between sensorsensitivity and loop interaction,a

    function was defined by the difference between the absolute

    values of the elements of the U vectors as:

    Zi = |U1i | |U2i |. (2)

    The maximum and the minimum of this function as sug-

    gested by Moore (1992) are selected as the two tray locations

    for the temperature control points. For an example, Fig. 10

    shows the Zi for the Inventory Strategy #1 with two manip-

    ulated variables of aqueous reflux and reboiler duty. From

    this figure, temperatures at stages #6 and #16 are selected

    for the two controlled variables for the above two manip-

    ulated variables. To compare among the alternative control

    structures, condition number (CN) and relative gain array

  • 7/27/2019 1-s2.0-S0009250904004282-main

    11/21

    I.L. Chien et al. / Chemical Engineering Science 59 (2004) 4547 4567 4557

    Table 11

    SVD and RGA analysis for each control structure

    Control Controller pairing Singular values CN RGA(11)

    structure

    CS1 (Inventory T16aqueous reflux 1 = 102.9 10.05 2.64

    Strategy #1) T6reboiler duty 2 = 10.24

    CS2 (Inventory T7organic reflux 1 = 112.2 3.14 0.673Strategy #1) T17reboiler duty 2 = 35.76

    CS3 (Inventory T16aqueous reflux 1 = 60.72 3.27 0.874

    Strategy #1) T7organic reflux 2 = 18.57

    CS4 (Inventory T6aqueous reflux 1 = 108.0 7.89 2.24

    Strategy #2) T15reboiler duty 2 = 13.68

    CS5 (Inventory T7entrainer makeup 1 = 101.6 529.2 0.594

    Strategy #2) T16reboiler duty 2 = 0.192

    CS6 (Inventory T16aqueous reflux 1 = 38.93 282.1 0.595

    Strategy #2) T6entrainer makeup 2 = 0.138

    (RGA) are also calculated for each control structure. The

    candidate control structures are listed below:

    CS1: using Inventory Strategy #1 with reboiler duty and

    aqueous reflux as two manipulated variables for dual-

    point temperature control.

    CS2: using Inventory Strategy #1 with reboiler duty and

    organic reflux as two manipulated variables for dual-

    point temperature control.

    CS3: using Inventory Strategy #1 with aqueous reflux and

    organic reflux as two manipulated variables for dual-

    point temperature control.

    CS4: using Inventory Strategy #2 with reboiler duty and

    aqueous reflux as two manipulated variables for dual-

    point temperature control.CS5: using Inventory Strategy #2 with reboiler duty and

    entrainer makeup as two manipulated variables for

    dual-point temperature control.

    CS6: using Inventory Strategy #2 with aqueous reflux and

    entrainer makeup as two manipulated variables for

    dual-point temperature control.

    Table 11 summarizes the results of SVD and RGA anal-

    ysis for the above six control structures. From the results of

    this table, several guidelines as suggested in Moore (1992)

    are followed to screen out the undesirable control structures

    from this steady-state analysis. The guidelines are: to select

    the smallest singular value as large as possible; to select

    the CN as small as possible; and to select the RGA(11) as

    close to unity as possible. From these guidelines, CS5 and

    CS6 are dropped for further comparison. For the remain-

    ing four control structures (CS14), further closed-loop dy-

    namic evaluation will be made to determine which control

    structure is the best.

    For the manipulated variables not used for temperature

    control purpose, it is preferable to design some kind of ratio

    scheme in order to move these manipulated variables ac-

    cording to the disturbance changes. For example, with con-

    trol structure CS1, constant ratio of organic reflux flow rate

    to feed flow rate is maintained throughout the closed-loop

    simulation run. In order to compensate the feed disturbance

    effect dynamically, a first-order lag with adjustable time con-

    stant is also included in the ratio scheme.

    Table 11 only shows the steady-state characteristics of

    each control structure. However, good steady-state charac-

    teristics are not a sufficient condition for good dynamic con-

    trol system performance. Thus, Aspen DynamicsTM will be

    used to evaluate the control system performance for the alter-

    native control structures. Since the 10% unmeasured feed

    composition changes are the more severe closed-loop test in

    comparison with the feed rate changes, these load changes

    will be made in the closed-loop dynamic simulations for

    comparison. All level loops are assumed to be controlled by

    P-only controller in order to smooth out their manipulatedvariables in the system. Controller gain of 2.0 as suggested

    in Luyben (2002) is used in all the level loops. The PID tun-

    ing constants for all the stage temperature control loops are

    tuned using the same multiloop tuning guideline (cf. Chien

    et al., 1999a), thus fair comparison can be made on the

    closed-loop dynamic responses among the four candidate

    control structures. Besides the dynamic response, the eval-

    uation of which control structure is adequate will emphasis

    more on the observation at final steady-state if the candidate

    control structure will actually meet the final product speci-

    fications in spite of the load disturbances.

    The closed-loop dynamic responses of control structures

    CS1CS4 with 10% changes in the feed H2O composition

    are shown in Figs. 1114, respectively. The disturbances are

    introduced at time = 0.5 h. With Inventory Strategy #1 (or-

    ganic phase level to manipulate the entrainer makeup flow),

    the maximum makeup flow is assumed to be larger than

    twice of the steady-state flow rate in order to have better con-

    trol of the organic phase level when this level is dropping.

    Since the control loop pairing of CS1 is unconventional (re-

    boiler duty to control temperature at a stage closer to the

    top of the column), it is very difficult to find proper tuning

    constants for the two temperature control loops. The tuning

    guideline in Chien et al. (1999a) has to be further detuned

  • 7/27/2019 1-s2.0-S0009250904004282-main

    12/21

    4558 I.L. Chien et al. / Chemical Engineering Science 59 (2004) 45474567

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    T

    (C)

    92

    94

    96

    98

    100

    102

    104

    106

    108

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    T(

    C)

    110

    112

    114

    116

    118

    120

    122

    124

    126

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    ReboilerDuty(KW)

    152

    154

    156

    158

    160

    162

    164

    166

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    AqueousReflux(mol/min)

    15

    20

    25

    30

    35

    40

    45

    50

    Feed H O +10% Change

    Feed H O -10% Change

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    OrganicLevel(m)

    0.40

    0.45

    0.50

    0.55

    0.60

    0.65

    0.70

    0.75

    0.80

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    MakeupFlow(mo

    l/min)

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    1.2

    1.4

    1.6

    1.8

    Feed H O +10% Change

    Feed H O -10% Change

    Fig. 11. Closed-loop dynamic simulation using control structure CS1 with 10% changes in the feed H2O composition.

    in order to make the system stabilized. Notice that for feed

    H2O 10% change in Fig. 11, entrainer makeup flow valve

    has to be fully open at time = 2.4 h and stayed fully open

    until time = 10.2 h. The controlled temperature points are

    not steady yet at final simulation time of 20 h. For CS2 in

    Fig. 12, the dynamic response of 10% feed H2O change

    is quite satisfactory. Two controlled temperature points are

    returned back to setpoints well before time = 20 h. How-

    ever, the dynamic response is unacceptable for +10% feed

    H2O change. Although two controlled temperature points

    reach their setpoints at final simulation time, the dynamic

    response of organic phase level is very bad. Its manipu-

    lated variable switches from fully close to fully open for

    the entire simulation run. Similar unacceptable dynamic re-

    sponses are observed in Fig. 13 for CS3. In comparison with

    the other three control structures, it is quite obvious from

    Fig. 14 that CS4 is the best control structure in terms of the

    dynamic response. All controlled and manipulated variables

    reach new steady-state within 8h. The organic phase level

    is also maintained much better than the other three control

    structures.

    Although not shown in this paper, dynamic closed-loop

    tests for CS5 and CS6 are also performed for 10% changes

    in the feed H2O composition. The closed-loop performance

  • 7/27/2019 1-s2.0-S0009250904004282-main

    13/21

    I.L. Chien et al. / Chemical Engineering Science 59 (2004) 4547 4567 4559

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    T

    (C)

    90

    92

    94

    96

    98

    100

    102

    104

    106

    108

    110

    112

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    T

    (C)

    119.0

    119.2

    119.4

    119.6

    119.8

    120.0

    120.2

    120.4

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    OrganicR

    eflux(mol/min)

    80

    90

    100

    110

    120

    130

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    ReboilerDuty(KW)

    140

    150

    160

    170

    180

    190

    200

    Feed H O +10% Change

    Feed H O -10% Change

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    OrganicLevel(m)

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    MakeupFlow

    (mol/min)

    -0.2

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    1.2

    1.4

    1.6

    1.8

    Feed H O +10% Change

    Feed H O -10% Change

    Fig. 12. Closed-loop dynamic simulation using control structure CS2 with 10% changes in the feed H2O composition.

    is also not satisfactory particularly for the temperature loop

    using entrainer makeup flow as the manipulated variable.

    This ineffectiveness of the entrainer makeup flow on the con-

    trolled tray temperature can actually be seen in previous sen-

    sitivity plot (Fig. 9). Large 10% changes in the entrainer

    makeup only give comparable tray temperature perturbations

    to 0.1% changes in the aqueous reflux flow. The large vari-

    ations on the entrainer makeup flow rate as well as the con-

    trolled tray temperature have detrimental effect on the prod-

    uct composition specifications. One thing worth mention is

    that although the closed-loop dynamic response of CS6 is

    much worse than CS4 but it is performed somewhat better

    than CS1, CS2, and CS3. This is another proof that good

    steady-state characteristics are not a sufficient condition for

    good dynamic control system performance of nonlinear sys-

    tems. The condition number of CS6 (282.1 in Table 11) is

    much larger than CS1CS3 but gives better dynamic control

    performance.

    The final objective of the alternative control structures

    is to maintain the bottom and top products specifications

    in spite of the load disturbances. Figs. 15 and 16 compare

    the dynamic responses of these four control structures with

    +10% and 10% feed H2O composition changes, respec-

    tively. It is obvious from these two figures that CS4 is the

    best control structure to reject feed H2O composition dis-

    turbances. Both bottom and top product compositions are

  • 7/27/2019 1-s2.0-S0009250904004282-main

    14/21

    4560 I.L. Chien et al. / Chemical Engineering Science 59 (2004) 45474567

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    T

    (C)

    90

    92

    94

    96

    98

    100

    102

    104

    106

    108

    110

    112

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    T

    (C)

    95

    100

    105

    110

    115

    120

    125

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    OrganicR

    eflux(mol/min)

    80

    90

    100

    110

    120

    130

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    AqueousReflux(mol/min)

    -10

    0

    10

    20

    30

    40

    50

    60

    Feed H O +10% Change

    Feed H O -10% Change

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 1 0 11 12 13 14 15 16 17 18 19 20

    OrganicLevel(m)

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    1.2

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    MakeupFlow(mol/min)

    -0.2

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    1.2

    1.4

    Feed H O +10% Change

    Feed H O -10% Change

    Fig. 13. Closed-loop dynamic simulation using control structure CS3 with 10% changes in the feed H2O composition.

    quickly returned back to specifications much faster than the

    other three alternative control structures. Control structure

    CS3 dynamically departs the furthest for both bottom and

    top product compositions to their specifications for +10%

    feed H2O composition change and control structure CS1

    dynamically departs the furthest for 10% feed H2O com-

    position change. In terms of the final steady-state value,

    control structure CS2 departs the furthest to bottom prod-

    uct specification (see Fig. 16). This control structure of CS2

    makes the aqueous reflux flow rate fixed during the dy-

    namic runs thus lose the ability to adjust the aqueous flow

    rate upward or downward to cope with the feed H2O com-

    position changes. Attempts have been made to introduce a

    ratio scheme to maintain constant aqueous reflux ratio in-

    stead of ratio to feed flow. The dynamic results are even

    worse than fixing the aqueous reflux flow rate during load

    disturbances.

    Another important closed-loop test for this control system

    is the feed flow rate changes. These changes are necessary

    in order to adjust the production rate upward or downward.

    Fig. 17 shows the dynamic responses for the proposed con-

    trol structure CS4 under 10% feed rate changes. Notice

    again that the closed-loop dynamic response is very sat-

    isfactory. Although not shown in the paper, both product

    specifications are maintained in spite of the production rate

    changes.

  • 7/27/2019 1-s2.0-S0009250904004282-main

    15/21

    I.L. Chien et al. / Chemical Engineering Science 59 (2004) 4547 4567 4561

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    T

    (C)

    95

    96

    97

    98

    99

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 1 0 11 12 13 14 15 16 17 18 19 20

    T

    (C

    )

    114

    115

    116

    117

    118

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 1 0 11 12 13 14 15 16 17 18 19 20

    AqueousR

    eflux(mol/min)

    20

    25

    30

    35

    40

    45

    50

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 1 0 11 12 13 14 15 16 17 18 19 20

    ReboilerDuty(KW)

    156

    158

    160

    162

    164

    166

    168

    170

    Feed H O +10% Change

    Feed H O -10% Change

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    OrganicLevel(m

    )

    0.718

    0.720

    0.722

    0.724

    0.726

    0.728

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    OrganicReflux(mol/min)

    88

    90

    92

    94

    96

    98

    Feed H O +10% Change

    Feed H O -10% Change

    Fig. 14. Closed-loop dynamic simulation using control structure CS4 with 10% changes in the feed H2O composition.

    3.1.1. Summary of dynamic simulation runs

    Some concluding remarks can be made after the above

    dynamic runs. Firstly, good steady-state characteristics on

    CS1CS3 (see Table 11) are not a sufficient condition for

    good dynamic control system performance of nonlinear

    systems. These three control structures gave much un-

    acceptable closed-loop performance than CS4. Secondly,

    dynamic simulations show that it is better to avoid using

    entrainer makeup flow as the manipulated variable for either

    of the organic phase level loop (CS1CS3) or the controlled

    tray temperature loop (CS5 and CS6). In this system, there

    are three free manipulated variables that can be cho-

    sen for the dual temperature loop. If including the organic

    phase level loop, there are total of four manipulated

    variables that can be chosen from. Thus, it is possible to

    select a control strategy not using entrainer makeup flow

    as manipulated variable. This entrainer makeup flow rate

    is fixed at the base case value and ratio to fresh feed rate

    changes.

    Fixing this entrainer makeup flow at base case value un-

    der various feed composition changes can still meet prod-

    uct purity specifications. This can be demonstrated by some

    steady-state simulation runs as in previous Table 9. In that ta-

    ble, desirable entrainer makeup flow rate is at 0.16mol/min

  • 7/27/2019 1-s2.0-S0009250904004282-main

    16/21

    4562 I.L. Chien et al. / Chemical Engineering Science 59 (2004) 45474567

    Time(hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    TopAceticAc

    idComposition

    0.00

    0.02

    0.04

    0.06

    0.08

    0.10

    0.12

    0.14

    0.16

    CS1

    CS2

    CS3

    CS4

    Time(hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    BottomAceticAcidCom

    position

    0.84

    0.86

    0.88

    0.90

    0.92

    0.94

    0.96

    0.98

    1.00

    1.02

    CS1

    CS2

    CS3

    CS4

    Fig. 15. Comparison of bottom product acetic acid compositions for the

    four control structures with +10% feed H2O composition change.

    for 50% feed H2O composition. If feed H2O composition

    is changed to 40%, from the table, the desirable entrainer

    makeup flow rate is at 0.13 mol/min. This does not mean that

    entrainer makeup flow rate has to be at this value to meet

    two product purity specifications. As mentioned previously,

    there are multiple steady-state conditions which can meet

    both product purity specifications because there are three

    degrees of freedom. In fact, a steady-state run at 40% feed

    H2O composition can be made with fixing of the two prod-

    uct purity specifications by varying reboiler duty and aque-

    ous reflux and also fixing the entrainer makeup flow rateat 0.16 mol/min. The resulting steady-state condition gave a

    little bit more on the value of the reboiler duty (176 KW vs.

    175KW in Table 9) but still holding two product purity spec-

    ifications. Also, noticeably the impurity of the bottom prod-

    uct shifted to more in iBuAc and less in H2O but the total

    impurity (iBuAc+H2O) is still at 0.1 mol% as desired. This

    demonstrates that control structure to fix entrainer makeup

    flow rate is workable.

    Since the existence of this extra degree of freedom (aque-

    ous reflux) to let the entrainer makeup flow released from

    controlling the organic phase level or tray temperature, it is

    desirable to have this extra degree of freedom in the sys-

    Time(hr)

    0 1 2 3 4 5 6 7 8 9 1 0 11 12 13 14 15 16 17 18 19 20

    TopAceticAcidComposition

    0.000

    0.005

    0.010

    0.015

    0.020CS1

    CS2

    CS3

    CS4

    Time(hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    BottomAceticAcidCom

    position

    0.970

    0.975

    0.980

    0.985

    0.990

    0.995

    1.000

    1.005

    CS1

    CS2

    CS3

    CS4

    Fig. 16. Comparison of bottom product acetic acid compositions for the

    four control structures with 10% feed H2O composition change.

    tem. This extra degree of freedom can provide desirable

    closed-loop dynamic response, thus if the feed composi-

    tion is rich in water (70mol% H2O), it is preferable to

    add a preconcentrator column before the feed stream to im-

    prove the dynamic behavior. This suggestion is supported

    from previous Fig. 7 and the dynamic simulation runs in this

    section.

    From the dynamic responses of CS4 (Fig. 14), another

    important observation can be made. With disturbances like

    10% changes in the feed H2O composition, the aqueous

    reflux flow rate will be adjusted upward or downward ac-

    cordingly in order to maintain about the same overall H2O

    composition into the column. However, the reboiler duty is

    actually returned back close to their original steady-state af-

    ter some dynamic transients. This inspires the thinking if

    simpler single temperature control strategy will work or not?

    The attempts of using single temperature control strategy

    will be explored next.

    3.2. Simpler single temperature loop control strategy

    The idea of the simpler single temperature loop control

    strategy is to use aqueous reflux flow rate to hold some tray

  • 7/27/2019 1-s2.0-S0009250904004282-main

    17/21

    I.L. Chien et al. / Chemical Engineering Science 59 (2004) 4547 4567 4563

    Time(hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    T(C)

    97

    Feed Rate +10% Change

    Feed Rate -10% Change

    Time(hr)

    0 1 2 3 4 5 6 7 8 9 1 0 11 12 13 14 15 16 17 18 19 20

    T

    (C

    )

    113

    114

    115

    116

    117

    118

    119

    120

    Time(hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    AqueousReflux(mol/min)

    30

    31

    32

    33

    34

    35

    36

    37

    38

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    ReboilerDuty

    (KW)

    140

    150

    160

    170

    180

    190

    Time(hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    OrganicLevel(m)

    0.710

    0.715

    0.720

    0.725

    0.730

    0.735

    Time(hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    OrganicReflux(mol/min)

    80

    82

    84

    86

    88

    90

    92

    94

    96

    98

    100

    102

    104

    Feed Rate +10% Change

    Feed Rate -10% Change

    Feed Rate +10% ChangeFeed Rate -10% Change

    Feed Rate +10% Change

    Feed Rate -10% Change

    Feed Rate +10% Change

    Feed Rate -10% Change

    Feed Rate +10% Change

    Feed Rate -10% Change

    Fig. 17. Closed-loop dynamic simulation using control structure CS4 with 10% changes in the feed flow rate.

    temperature inside the column and to keep the other two

    free manipulated variables (reboiler duty and entrainer

    makeup) to maintain ratioed to the feed flow rate. From the

    earlier sensitivity analysis in Fig. 9, the most sensitive con-

    trol point inside the column is stage #6 which is closer to

    the top of the column. Since the main acetic acid product is

    drawn from the bottom of the column, an alternative con-

    trol point is to select the second most sensitive control point

    which is closer to the bottom of the column. This alterna-

    tive control point will be stage #16. Thus two single loop

    control structures will be deduced for closed-loop dynamic

    test. They are:

    CS7: using Inventory Strategy #2 in the previous subsection

    with temperature at stage #6 controlled by manipulat-

    ing aqueous reflux flow, while the other two manip-

    ulated variables, reboiler duty and entrainer makeup,

    maintain constant ratios to the feed flow rate.

    CS8: using Inventory Strategy #2 in the previous subsec-

    tion with temperature at stage #16 controlled by ma-

    nipulating aqueous reflux flow, while the other two

    manipulated variables, reboiler duty and entrainer

  • 7/27/2019 1-s2.0-S0009250904004282-main

    18/21

    4564 I.L. Chien et al. / Chemical Engineering Science 59 (2004) 45474567

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    T

    (C)

    94

    95

    96

    97

    98

    99

    100

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 1 0 11 12 13 14 15 16 17 18 19 20

    AqueousReflux(mol/min)

    20

    25

    30

    35

    40

    45

    50

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    OrganicLevel(m)

    0.7215

    0.7220

    0.7225

    0.7230

    0.7235

    0.7240

    0.7245

    0.7250

    0.7255

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    OrganicReflux(mol/min)

    90.4

    90.6

    90.8

    91.0

    91.2

    91.4

    91.6

    91.8

    92.0

    92.2

    92.4

    92.6

    92.8

    93.0

    93.2

    93.4

    93.6

    Feed H O +10% Change

    Feed H O -10% Change

    Feed H O +10% Change

    Feed H O -10% Change

    Fig. 18. Closed-loop dynamic simulation using control structure CS7 with 10% changes in the feed H2O composition.

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 1 0 11 12 13 14 15 16 17 18 19 20

    T

    (C)

    112

    114

    116

    118

    120

    122

    124

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 1 0 11 12 13 14 15 16 17 18 19 20

    AqueousReflux(mol/min)

    15

    20

    25

    30

    35

    40

    45

    50

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 1 0 11 12 13 14 15 16 17 18 19 20

    OrganicLevel(m)

    0.7220

    0.7225

    0.7230

    0.7235

    0.7240

    0.7245

    0.7250

    Feed H O +10% Change

    Feed H O -10% Change

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 1 0 11 12 13 14 15 16 17 18 19 20

    OrganicReflux(mol/min)

    90.6

    90.8

    91.0

    91.2

    91.4

    91.6

    91.8

    92.0

    92.2

    92.4

    92.6

    92.8

    93.0

    93.2 Feed H O +10% Change

    Feed H O -10% Change

    Feed H O +10% Change

    Feed H O -10% Change

    Fig. 19. Closed-loop dynamic simulation using control structure CS8 with 10% changes in the feed H2O composition.

  • 7/27/2019 1-s2.0-S0009250904004282-main

    19/21

    I.L. Chien et al. / Chemical Engineering Science 59 (2004) 4547 4567 4565

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 1 0 11 12 13 14 15 16 17 18 19 20

    TopAceticAc

    idComposition

    0.0007

    0.0008

    0.0009

    0.0010

    0.0011

    0.0012

    0.0013

    CS7

    CS8

    CS4

    Time (hr)0 1 2 3 4 5 6 7 8 9 1 0 11 12 13 14 15 16 17 18 19 20

    BottomAceticAcidCo

    mposition

    0.9978

    0.9980

    0.9982

    0.9984

    0.9986

    0.9988

    0.9990

    0.9992

    0.9994

    CS7CS8

    CS4

    Fig. 20. Comparison of bottom product acetic acid compositions for the

    double loop control structure of CS4 with single loop control structures

    of CS7 and CS8 with +10% feed H2O composition change.

    makeup, maintain constant ratios to the feed flow

    rate.

    Figs. 18 and 19 show the closed-loop dynamic responses

    for CS7 and CS8 under 10% feed H2O composition

    changes, respectively. Notice that the dynamic responses

    are all quite satisfactory with all variables settled out at

    new steady-state values even faster than CS4 (comparing to

    Fig. 14). The dynamic responses of the most important bot-

    tom and top product compositions are shown in Figs. 20 and

    21 for +10% and

    10% changes in the feed H2O compo-sition, respectively. Notice first that the scaling ofFigs. 20

    and 21 are much smaller than previous Figs. 15 and 16

    indicating these two single loop control structures perform

    much better than previous CS1, CS2, and CS3. Comparing

    to more complex double loop control structure CS4, CS7

    performs very satisfactory. Both product compositions are

    maintained at tight specifications even the control point is

    far away from the column bottom. On the contrary, the

    acetic acid loss through the column top cannot be main-

    tained at tight specification for control structure CS8. The

    proposed control structure CS7 also performs very well for

    10% feed rate disturbances. The dynamic responses for

    Time(hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    TopAceticAc

    idComposition

    0.000

    0.001

    0.002

    0.003

    0.004

    0.005

    0.006

    CS7CS8

    CS4

    Time(hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    BottomAceticAcidComposition

    0.9972

    0.9974

    0.9976

    0.9978

    0.9980

    0.9982

    0.9984

    0.9986

    0.9988

    0.9990

    CS7

    CS8

    CS4

    Fig. 21. Comparison of bottom product acetic acid compositions for the

    double loop control structure of CS4 with single loop control structures

    of CS7 and CS8 with 10% feed H2O composition change.

    the temperature loop and the organic level loop are shown in

    Fig. 22. All controlled and manipulated variables reach new

    steady-state values after a short dynamic transient.Although

    not shown in the paper, both product compositions are also

    maintained at tight specifications. The final proposed simple

    single temperature loop control structure of CS7 is shown in

    Fig. 23. Although the two product purities are assumed not

    to be measured on-line, but if they can be measured in-

    frequently in quality lab, small trimming of the controlled

    temperature setpoint or small changes of the reboiler heat

    duty can be made to even more precisely to hold the prod-uct purities at their specifications during sustained feed

    disturbances.

    4. Conclusions

    Three candidate entrainers (ethyl acetate, iso-butyl ac-

    etate, and n-butyl acetate) are considered for acetic acid

    dehydration via heterogeneous azeotriopic distillation. The

    factors needed to be considered in selecting the proper en-

    trainer are illustrated for this example system. Optimum col-

    umn designs and operating conditions are obtained for these

  • 7/27/2019 1-s2.0-S0009250904004282-main

    20/21

    4566 I.L. Chien et al. / Chemical Engineering Science 59 (2004) 45474567

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 1 0 11 12 13 14 15 16 17 18 19 20

    T

    (C)

    94

    96

    98

    100

    102

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    AqueousR

    eflux(mol/min)

    25

    30

    35

    40

    45

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 1 0 11 12 13 14 15 16 17 18 19 20

    OrganicLevel(m)

    0.71

    0.72

    0.73

    0.74

    Time (hr)

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    OrganicReflux

    (mol/min)

    75

    80

    85

    90

    95

    100

    105

    110

    Feed Rate +10% Change

    Feed Rate -10% Change

    Feed Rate +10% Change

    Feed Rate -10% Change

    Feed Rate +10% Change

    Feed Rate -10% Change

    Feed Rate +10% Change

    Feed Rate -10% Change

    Fig. 22. Closed-loop dynamic simulation using control structure CS7 with 10% changes in the feed flow rate.

    Fig. 23. Schematic diagram of the proposed control structure CS7.

  • 7/27/2019 1-s2.0-S0009250904004282-main

    21/21

    I.L. Chien et al. / Chemical Engineering Science 59 (2004) 4547 4567 4567

    three candidate systems using rigorous process simulation.

    Total Annual Cost (TAC) is used as the objective function in

    determining the optimum column designs andoperating con-

    ditions for these three candidate systems. Iso-butyl acetate

    was found to be the best entrainer with resulting TAC only

    about 55% of the system with no entrainer. The optimum

    overall control strategy is also proposed for this column sys-tem to hold both bottom and top product specifications in

    spite of10% feed rate and 10% feed H2O composition

    load disturbances. Several alternative control structures are

    compared using dynamic simulation. The proposed overall

    control strategy is very simple requiring only one tray tem-

    perature control loop inside the column. This simple overall

    control strategy can easily be implemented in industry for

    wider applications.

    Acknowledgements

    This work is supported by National Science Council ofR. O. C. under grant nos. NSC 89-2214-E-011-025 and NSC

    90-2214-E-011-013. Helpful suggestions from anonymous

    reviewers are gratefully acknowledged.

    References

    Aspen Technology, Inc., 2001. Aspen Plus Users Manual 11.1. Aspen

    Technology, Inc., Cambridge.

    Bozenhardt, H.F., 1988. Modern control tricks solve distillation problems.

    Hydrocarbon Processing 67 (6), 4750.

    Chiang, S.F., Kuo, C.L., Yu, C.C., Wong, D.S.H., 2002. Design alternatives

    for the amyl acetate process: coupled reactor/column and reactive

    distillation. Industrial and Engineering Chemistry Research 41 (13),

    32333246.

    Chien, I.L., Huang, H.P., Yang, J.C., 1999a. A simple multiloop tuning

    method for PID controllers with no proportional kick. Industrial and

    Engineering Chemistry Research 38 (4), 14561468.

    Chien, I.L., Wang, C.J., Wong, D.S.H., 1999b. Dynamics and control of

    a heterogeneous azeotropic distillation column: conventional control

    approach. Industrial and Engineering Chemistry Research 38 (2),

    468478.

    Chien, I.L., Wang, C.J., Wong, D.S.H., Lee, C.H., Cheng, S.H., Shih, R.F.,

    Liu, W.T., Tsai, C.S., 2000a. Experimental investigation of conventional

    control strategies for a heterogeneous azeotropic distillation column.

    Journal of Process Control 10 (4), 333340.

    Chien, I.L., Chen, W.H., Chang, T.S., 2000b. Operation and decoupling

    control of a heterogeneous azeotropic distillation column. Computers

    and Chemical Engineering 24 (27), 893899.

    Christensen, S.P., Olson, J.D., 1992. Phase equilibria and multiple

    azeotrope of the acetic acidisobutyl acetate system. Fluid Phase

    Equilibria 79, 187199.

    Costantini, G., Serafini, M., Paoli, P., 1981. Process for the recovery of

    the solvent and of the by-produced methylacetate in the synthesis of

    terephthalic acid. U.S. Patent 4, 250, 330.

    Douglas, J.M., 1988. Conceptual Process Design. McGraw-Hill, New

    York.

    Gaubert, M.A., Gerbaud, V., Joulia, X., Peyrigain, P.S., Pons, M., 2001.

    Analysis and multiple steady states of an industrial heterogeneous

    azeotropic distillation. Industrial and Engineering Chemistry Research

    40 (13), 29142924.

    Gmehling, J., Onken, U., 1977. Vaporliquid equilibrium data collection.

    In: Behrens, D., Eckermann, R. (Eds.), DECHEMA Chemistry Data

    Series. DECHEMA Publishers, Frankfurt, Germany.

    Hayden, J.G., OConnell, J.P., 1975. A generalized method for predicting

    second virial coefficients. Industrial and Engineering Chemistry Process

    Design and Development 14, 209.

    Horsley, L.H., 1973. Azeotropic Data III. Advances in Chemistry Series

    No. 116. American Chemical Society, Washington, DC.Kurooka, T., Yamashita, Y., Nishitani, H., Hashimoto, Y., Yoshida, M.,

    Numata, M., 2000. Dynamic simulation and nonlinear control system

    design of a heterogeneous azeotropic distillation column. Computers

    and Chemical Engineering 24 (27), 887892.

    Luyben, M.L., Tyreus, B.D., 1998. An industrial design/control study

    for the vinyl acetate monomer process. Computers and Chemical

    Engineering 22 (78), 867877.

    Luyben, W.L., 2002. Plantwide Dynamic Simulators in Chemical

    Processing and Control. Marcel Dekker, New York.

    Moore, C.F., 1992. Selection of controlled and manipulated variables.

    In: Luyben, W.L. (Ed.), Practical Distillation Control. Van Nostrand

    Reinhold, New York.

    Mitsui Petro-Chemical Industries, Ltd., 1980. Process for azeotropic

    distillation. United Kingdom Patent 1, 576, 787.

    Othmer, D.F., 1936. Process for dehydration of acetic acid and otherlower fatty acids. U.S. Patent 2, 050, 234.

    Othmer, D.F., 1941. Azeotropic distillation for dehydrating acetic acid.

    Chemical and Metallurgical Engineering 40, 9195.

    Othmer, D.F., 1963. Azeotropic separation. Chemical Engineering Progress

    59 (6), 6778.

    Parten, W.D., Ure, A.M., 1999. Dehydration of acetic acid by azeotropic

    distillation in the production of an aromatic acid. U.S. Patent 5, 980,

    696.

    Pham, H.N., Doherty, M.F., 1990. Design and synthesis of heterogeneous

    azeotropic distillations-III. column sequences. Chemical Engineering

    Science 45 (7), 18451854.

    Renon, H., Prausnitz, J.M., 1968. Local compositions in thermodynamics

    excess functions for liquid mixtures. A.I.Ch.E. Journal 14, 135.

    Rovaglio, M., Faravelli, T., Biardi, G., Gaffuri, P., Soccol, S., 1993.

    The key role of entrainer inventory for operation and control ofheterogeneous azeotropic distillation column towers. Computers and

    Chemical Engineering 17 (5), 535.

    Siirola, J.J., 1995. An industrial perspective on process synthesis. In:

    Biegler, L.T., Doherty, M.F. (Eds.), A.I.Ch.E. Symposium Series No.

    304, vol. 91, 222233.

    SZrensen, J.M., Arlt, W., 1979. Liquidliquid equilibrium data collection

    binary systems. In: Behrens, D., Eckermann, R. (Eds.), DECHEMA

    Chemistry Data Series. DECHEMA Publishers, Frankfurt, Germany.

    SZrensen, J.M., Arlt, W., 1980. Liquidliquid equilibrium data collection

    ternary systems. In: Behrens, D., Eckermann, R. (Eds.), DECHEMA

    Chemistry Data Series. DECHEMA Publishers, Frankfurt, Germany.

    Tanaka, S., Yamada, J., 1972. Graphical calculation method for minimum

    reflux ratio in azeotropic distillation. Journal of Chemical Engineering

    of Japan 5, 2026.

    Ulrich, J., Morari, M., 2002. Influence of impurities on the controlof heterogeneous azeotropic distillation columns. Industrial and

    Engineering Chemistry Research 41 (2), 230250.

    Wasylkiewicz, S.K., Kobylka, L.C., Castillo, F.J.L., 2000. Optimal design

    of complex azeotropic distillation columns. Chemical Engineering

    Journal 79, 219227.

    Widagdo, S., Seider, W.D., 1996. Azeotropic distillation. A.I.Ch.E. Journal

    42 (1), 96130.