Nonlinear Model Predictive Control of A

Embed Size (px)

Citation preview

  • 7/24/2019 Nonlinear Model Predictive Control of A

    1/20

    Compurers them. Engng,

    Vol. 18, No.

    2, PP. 83-102, 1994

    0098-I 354/94 6.00

    + 0.00

    Printed in Great Britain. All rights reserved Copyright 0 1994 Pergamon Press

    Ltd

    NONLINEAR MODEL PREDICTIVE CONTROL OF A

    FIXED-BED WATER-GAS SHIFT REACTOR:

    AN EXPERIMENTAL STUDY

    G .

    T.

    WRIGHT

    and T. F.

    EDGAR

    Department of Chemical Engineering, The University of Texas at Austin, Austin, TX 78712, U.S.A.

    Rec ei v e d 16 No vember f 9 9 . 7 ; i n a l r e v i si o n r e c ei v e d 2 J u n e 1 993 ; r e c ei v e d f o r p u b l i c a t i o n I 6 J u ne 199 3 )

    Abstract-This paper describes new results on the experimental application of nonlinear model-predictive

    control (NMPC) to a fixed-bed water-gas shift (WGS) reactor. The development and experimental

    validation of an appropriate first-principles WGS reactor model,

    and how it impacts controller

    performance is discussed. The implementation of NMPC is computationally intense, requiring that a large

    nonlinear program (NLP) be

    solved at each sampling period. The significant computational burden

    dictates that a relatively slow sam pling rate be used. Infrequent sampling, however, diminishes disturbance

    rejection capabilities. To combat this problem, NM PC was implemented

    in a master-slave cascade

    configuration where a low-level liner controller, having a significantly faster sampling rate, was employed.

    The control study was performed using a PC-based distributed control system (DCS) One of the three

    processors was dedicated to NMPC calculations. A complete and rigorous implementation strategy is

    described in the paper, and the performance of NMPC for set-point tracking of this nonlinear process

    is shown to be superior to adaptive or linear control. We also illustrate the ease with which NMPC

    accommodated feedforward control.

    1. INTRODUCI-ION

    The sev erity of the nonlinearities in chemical pro-

    cesses influences the selection of control algorithms

    for successful control of a process. The linearization

    of nonlinear physical m odels about a nominal operat-

    ing point is a standard approach for control system

    design. This approach is adequate when the process

    nonlinearities are mild or when plant operation

    is constrained to a narrow region, but for highly

    nonlinear systems such as nonisothermal chemical

    reactors, linear controllers designed in this way

    may perform poorly. The wide range of operating

    conditions encountered in start-up or shut-down

    of continuous processs and trajectory tracking

    of batch processes also prove difficult for linear

    control.

    One technique that attempts to compensate for the

    inadequacies associated with linear controllers uses

    an adaptive feedback control law based upon current

    and past operating conditions. In general, an adap-

    tive controller can be thought of as a nonlinear

    control technique that has two distinctly different

    types of time-dependent variables operating on very

    different time scales. Presum ably, the parameters of

    the process m odel vary slowly with time, whereas the

    process states (e.g. temperature or concentration)

    change at a much faster rate. While many successful

    applications of adaptive control have been reported

    in the literature, successful control with an adaptive

    algorithm may be difficult to achieve. Because of this

    deficiency, there appears to be a need for further

    improvemen ts in nonlinear controllers, which is the

    motivation for this research.

    In recent years, model-based control strategies

    employing differential geometric theory have been

    used to effect linearization via state or output feed-

    back. Global feedback linearization involves trans-

    forming the states of a nonlinear system into an

    eqivalent system such that, under state or output

    feedback, the resulting dynamic system becomes ex-

    actly linear in a specified manner. Brockett (1978) was

    the first to employ this methodology, which was

    further advanced by Hunt ef al. (1983). While this

    technique is intuitively appealing, the necessary trans-

    formations required for implementation often fail to

    exist, or do not yield the desired level of robustness

    to model error. H ence, we seek a more broadly

    applicable nonlinear control strategy.

    Recent improveme nts in computer capabilities

    permit the use of rigorous models for real-time

    optimization and control. A control strategy which

    takes advantage of the increased power and speed

    of computers is nonlinear model-predictive control

    (NMPC). The predictive control strategy involves

    a repeated optimization of an open-loop performance

    objective over a finite horizon extending from the

    current time into the future as done for linear model-

    predictive control (Cutler and Ramaker, 1980; Clarke

    et al., 1987; Eaton and Rawlings, 1991). Ra wlings

    83

  • 7/24/2019 Nonlinear Model Predictive Control of A

    2/20

    and Muske (1991) have developed a nominally stabi-

    Iron oxide, copper-zinc oxide and cobah-

    lizing constrained model predictive controller using

    molybdenum catalysts are among the catalysts used

    an infinite prediction horizon. Linear models with

    to promote shift conversion. Iron oxide catalysts have

    equality and inequality constraints permit the optim-

    been studied in much greater detail than the others.

    ization problem to be solved directly using quadratic

    They are classified as high-temperature shift catalysts

    programming (QP). NMPC extends MPC to nonlin- with maximum operating temperatures of approx.

    ear systems and incorporates constraints in an ex-

    500C. Since our principal goal w as not to study the

    plicit manner. WGS reaction, but to use it as a model chemistry to

    The primary objective of this research was to better implement nonlinear control strategies, a well-

    develop an advanced nonlinear control strategy for a

    understood, high-temperature, iron-oxide shift cata-

    fixed-bed reactor and to apply the strategy exper- lyst (United Catalyst Cl2-3-05) was used. Previous

    imentally. Section 2 describes the construction of a work on a similar reactor by Bell and Edgar (1991)

    laboratory scale fixed-bed water-gas shift reactor

    used a sulfur-tolerant cobalt-molybdenum catalyst

    with computer facilities capable of acquiring data and that is much m ore difficult to characterize.

    implementing advanced control. In Sections 3, 4 and The experimental facility for the WGS reactor

    5, a first-principles model for the WGS reactor suit- (Fig. 1) consisted of five parts: the dry gas feed

    able for use in real-time, model-based control strat- processing system, the steam generator, the wet gas

    egies is developed. A solution technique is adopted,

    feed processing system, the fixed-bed reactor an d the

    model parameters are estimated and the reactor effluent gas processing system.

    model is validated. Section 6 is devoted to NMPC

    development and implementation issues, and in Sec-

    2. I.

    Dry gas feed pr ocess i ng

    tion 7, NMPC is evaluated experimentally and com- Carbon monoxide, carbon dioxide, hydrogen and

    pared to more traditional control strategies.

    nitrogen were supplied to the reactor from pressure-

    regulated cylinders via mass flow controllers (MFC).

    2. THE EXPERIMENTAL FACILITY AND OPERATING

    Compressed air was obtained from building header.

    PROCEDURE

    MFC isolation valves in conjunction with manual

    The water-gas shift reaction (WCS) arises in the

    bypass valves were used to direct the supply gases as

    production of ammonia, hydrogen and organic

    desired.

    chemicals. The reaction is reversible and mildly

    exothermic:

    2.2. S t eam gene r a t o r

    The steam generating system w as similar to a

    CO(g) + H@(g) - CQl(g) + H,(g)

    generator used by Bell and Edgar (1991). The steam

    AH ,, = -9.8

    kcal/mol.

    generator was designed to vaporize a known quantity

    of water and to superheat it enough to avoid conden-

    In typical industrial applications the dry process

    sation prior to entering the reactor. Deionized water,

    feed contains carbon monoxide, carbon dioxide,

    obtained from a building tap, was stored in 20 I

    hydrogen, small quantities of hydrocarbons and

    polyethylene reservoir. A Chem-Tech diaphragm

    sulfur impurities. A nominal dry-gas reactor feed

    pump was used to regulate flow from the reservoir.

    composition assuming insignificant quantities of

    The deionized water was vaporized in two-parallel

    hydrocarbon, is 40 CO, 40 CO, and 20 Hr.

    3 m sections of 16 in., 3 16 stainless-steel. The parallel

    The steam to dry-gas ratio m ay vary, but typically

    tubing wa s wrapped around a 750 W cartridge heater.

    satisfies the condition that the steam to CO ratio

    This assembly was housed in a 30.5 cm long, 1.25 in.

    exceeds four depending upon effluent composition

    dia. tube. The tubing w as packed with magnesium

    goals.

    oxide, an electrical insulator with high thermal con-

    The W GS reaction is run in either a single adia-

    ductivity. A K-type thermocouple was also placed in

    batic fixed-bed reactor or in multiple reactors in series the assembly to monitor the cartridge heater tempera-

    when high conversion of carbon monoxide is re-

    ture. The heater temperature was varied by manipu-

    quired. Steam quench streams are often located be-

    lating the power input to the generator. The exterior

    tween beds in a multi-reactor configuration. In either

    of the steam generator was insulated with 2 in. min-

    configuration bed behavior is primarily influenced by eral wool insulation.

    varying the reactor inlet temperature or the steam to

    dry-gas ratio. Disturbances may include fluctuations

    2.3. Wet -gas feed pr ocessin g

    in the dry-gas feed composition and flow, and up-

    After mixing the dry gases with the steam, the wet

    stream temperature variations that cannot be ade-

    gas mixture was heated to a temperature of approx.

    quately rejected.

    165C. This temperature was maintained using PID

    84

    G. T. WRIGHT and T. F. EDGAR

  • 7/24/2019 Nonlinear Model Predictive Control of A

    3/20

    Nonlinear model predictive control

    k

    W8tW

    Fig. 1. Schematic of the experimental facility for the water-gas shift reactor.

    control to mitigate upstream temperature fluctu-

    ations caused by the steam generator. Approximately

    0.75 m of 0.25 in. tubing ran from the steam/dry-gas

    mixing tee to the reactor inlet. A heat exchanger was

    constructed by wrapping this tubing with 3.65 m of

    ceramic fiber-insulated, 1.9 n/f nichrom e wire. Tem-

    perature at the reactor inlet was controlied by varying

    power to the resistance heater. This assembly com-

    prised the feed preheater.

    2.4.

    The r e a c t o r

    The body of the reactor w as positioned vertically

    and was constructed using 3.175 cm (1.25 in.), 321

    stainless-steel with a wall thickness of 0.089 cm

    (0.035 in.). The total reactor length w as 99.06 cm

    (39 in.). A pair of temperature measurem ents were

    taken axially at 11 equally-spaced points along the

    reactor. Each pair of measurem ents consisted of a

    centerline-bed temperature and a corresponding wall

    temperature. The first pair of thermocouples were

    located 11.43 cm (4.5 in.) from the top of the reactor

    and the last at 72.39 cm (28.5 in.). In order to mini-

    mize heat loss, two independent resistance heaters

    were wrapped tightly around the reactor. T he first

    heater extended from the inlet header to the fourth

    thermocouple pair. The second continued from there

    and extended to the last thermocouple pair. The

    reactor was then insulated with several layers of thick

    Cerablanket Insulation.

    The feed gases and the reactor inlet section were

    heated u sing a 300 W cartridge heater, housed in a

    thermo-well assembly located at the top of the reac-

    tor. Power to the cartridge heater was supplied in a

    mann er analogous to the steam generator heater. The

    volume su rrounding the thermo-well and extending

    to the first pair of thermocouples was packed with

    Pyrex glass beads. Th ese facilitated flow distribution

    prior to gas entry into the active bed. They also

    served as catalyst su pport when loading the reactor

    bed.

    The active bed of the reactor w as 36.6 cm long and

    extended from the first thermocouple pair through

    the seventh. The active bed was packed with iron-

    oxide catalyst pellets (0.3175 cm/O.125 in.), yielding a

    reactor diameter to particle diameter ratio of approx.

    9.4. T he void fraction in the catalyst bed was 0.35.

    The exit section of the reactor was packed with

    kaolin clay spheres. They supported the catalyst bed

    during norm al operation and rested upon a perfo-

    rated plate located at the reactor exit. Four thermo-

    couple pairs extended into the reactor exit section.

    Some were used to examine the validity of model

    boundary conditions discussed later. The total exit

    section was 51.05 cm (20.1 in.) in length.

  • 7/24/2019 Nonlinear Model Predictive Control of A

    4/20

    86

    G.

    T.

    WRIGHT and

    T. F.

    EDGAR

    2.5 . Ef l uent gus pr ocessin g

    Upon exiting the reactor, the product gases passed

    through a series of units designed to eliminate the

    water vapor content of the gas and to determine the

    effluent dry-gas composition. Infrared analyzers were

    used to determine effluent CO and CO2 compositions.

    This information coupled with inlet mass flowrates

    for each species was sufficient to completely deter-

    mine the exit composition by material balance.

    2.6.

    D i g i t a l equ i pmen t

    A global bus architecture, distributed control

    system DC S employing two 33 MHz Intel 32-bit

    80386/80387 IBM-PC compatible computers and one

    16 MHz Intel 16-bit 80386/80387 IBM-PC compat-

    ible was used in this research. This architecture

    functionally divides the tasks traditionally im-

    plemented by a single host computer among several

    autonomous computers. The 16 MHz machine w as

    used for primary data collection and storage, for

    trending and as an operator interface. The 33 MHz

    machines were used as platforms for highly intensive

    numerical computations such as the optimizations

    necessary of NMPC calculations. Ethernet adaptor-

    boards were installed in each of the computers to

    provide for communication. Thin coaxial Ethernet

    cable was used to wire the computers together. The

    software used to drive the DCS for the WGS reaction

    facility was Intellutions FIX DMACS, a commer-

    cially available software package available on several

    platforms. DMAC S has an open architecture allow-

    ing easy database access for user-written code. Sev-

    eral mechanisms were provided for executing user

    tasks including event-based execution, fixed interval

    execution, specific time execution and continuous

    execution. The second of these techniques was used

    predominantly and was implemented using a sched-

    uler which was designed to spawn independent pro-

    grams as user-specified intervals.

    For data acquisition and control, digital and anlog

    Optomux stations were employed. Each Optomux

    station consisted of a brain board and a mounting

    rack. The Optomux brain board is a controller which

    operates as a slave device to a host computer. The

    brain boards are, in essence, intelligent multiplexers,

    capable of accomplishing most tasks independent of

    the host computer system. The series of Optomux

    stations communicated with the host computer over

    an RS-422/485 serial link. The Optomux brain boards

    were configured for multidrop operation.

    2.7. Reac to r ope ra t i on

    The procedures used to treat the catalysts were

    those suggested by the catalyst manu facturer, United

    Catalysts, Inc. (UCI). Activation of the UC1 iron

    oxide Cl2 catalyst basically entailed employing a gas

    medium to reduce the catalyst from its oxidized state.

    The first step in the activation procedure was to

    prepare the catalyst by heating it from ambient to

    150C using a 2: 3 nitrogen to steam gas mixture.

    Steam flowrate was typically 6-8 SLM. Upon reach-

    ing this temperature activation was initiated by intro-

    ducing a wet process gas mixture with a 1:2 dry-gas

    to steam ratio. The dry-gas mixture consisted of a

    2: 2: I mixture of COI, Hz, CO. Steam flowrate was

    maintained at 68 SLM. The wet process gas was

    used to raise the bed temperature to reaction con-

    ditions (300C). Special care was exercised through-

    out the heating and reducing phases to limit the rate

    of temperature increase in the bed to 65C. Th e inlet

    temperature was adjusted downward as reaction

    slowly proceeded to make certain that catalyst bed

    temperatures never exceeded approx. 465C. When

    the temperatures ultimately stabilized, the reducing

    procedure was complete. The entire process typically

    took from 15-20 h.

    Normal operation of the WGS reaction facility is

    categorized by start-up, data collection and control

    and shut-down. Open- and closed-loop experiments

    were performed at inlet temperatures ranging from

    270 to 300C. Total gas flowrates ranged from 10 to

    16 SLM. Steady-state experiments were performed by

    maintaining the active-bed inlet temperature at a

    desired value using PID control. This proved to be an

    effective method for rejecting upstream disturbances

    and replicating experiments. Closed-loop experiments

    were similarly performed using a cascade control

    scheme. An in depth discussion of this is given later

    in the text.

    3. DYNAMIC MODELING OF THE FIXED-BED WATER GAS

    SHIFT REACTOR

    When modeling fixed-bed catalytic reactors, con-

    sideration must be given to a multitude of phenom-

    ena ranging from fluid how to intra-particle and

    interphase mass and energy transport. These con-

    siderations lead to complex partial differential

    equation models, even for simple reaction schemes.

    The level of model detail is ultimately constrained by

    the availability of reliable physical property infor-

    mation for reactants and products, accurate rate

    expressions and knowledge of catalyst characteristics,

    among other things.

    How the model is to be used is another key

    constraint in the modeling effort. A rigorous fixed-

    bed reactor model would prove to be not only

    complex, but also unsuitable for on-line app lications

    such as optimization and nonlinear control. What

  • 7/24/2019 Nonlinear Model Predictive Control of A

    5/20

    Nonhnear

    mode l predictivecontrol 87

    follows is a discussion of the assumptions used to behave ideally. In addition, pressure drop across the

    develop a sufficiently simple, yet accurate, model for reactor was small, obviating the need for equations

    real-time implementation.

    describing this effect. The residence time

    for the

    The WG S reactor constructed for this study was process gas was on the order of 1 s. This was small

    designed to operate under nearly a diabatic con- compared to the time-scale for changes in catalyst-

    ditions. This wa s achieved by adding guard heaters to

    bed temperatures. Therefore, the quasi-steady state

    the reactor to minimize radial heat loss. The reactor

    assumption that concentration ch anges are instan-

    was also insulated w ith several inches of fiberglas taneous relative to temperature changes was adopted.

    insulation. While these efforts did not eliminate heat Dispersion is negligible when the ratio of reaction

    loss entirely, an adiabatic assum ption was employed length to particle diameter exceeds 100 (Carberry and

    for model simplification. Wendel, 1963, Rase, 1977). It was included here for

    Although fixed-bed reactors are heterogeneous sys-

    numerical conditioning as suggested by Windes

    terns with both fluid and solid phases, it is often

    (1986). The final model co nsisted of two partial

    reasonable to assume that the mass within a volume

    differential equations with axial and tempora l d imen-

    element can be characterized by a single bulk tem- sions. Danckwerts (1953) boundary conditions were

    perature, pressure and composition. The pseudo -

    used at the reactor inlet for the catalyst bed balances,

    homo geneous assumption is a valid approximation

    and zero gradient boundary conditions were applied

    provided comp osition and temperature gradients be-

    at the exit. The dimension less model is presented

    tween the fluid and solid phases are small. This

    below. Model symbo ls a re defined in the Nome ncla-

    situation prevails when reaction resistance is large

    ture.

    relative to mass and heat transfer resistance. Windes

    e? nf. (I 989) compared one- and tw o-phase models for

    3.1. Carbon monox i de ba l ance

    the oxidation of methanol. They concluded that

    qualitatively these mode ls compare favorably und er

    most circumstances.

    They further concluded that

    0= - +i [+]-Da (-i ,,) (1 )

    pe,

    even if the pseudo-hom ogeneous assump tion were

    3.2. Cata l ys t bed energy ba lan ce

    not strictly applicable,

    the one- and two-phase

    z-

    models compare well quantitatively with some par-

    Le

    af a?=

    ameter adjustment. Bell and Edgar (1991) employed

    x=z+& =

    I 1

    df2

    this assumption in a WGS reactor mode l for a system

    similar to the one constructed for this research. Their

    - %w@ - fw ) + fit-fco).

    (2)

    results confirmed that assuming homo geneity is a

    The WGS reactor model was nondimensiona lized

    practical simplification yielding good results for the

    using the following definitions:

    WGS reaction. The work of Ampay a and Rinker

    (1977) and Bonvin (1980) further supported this

    conclusion.

    f= ,

    rsf

    The spatial dimensionality of a fixed-bed reactor

    model may p rofoundly affect the models capacity for

    i=Z

    L

    accurate prediction. For small diameter reactors run-

    ning under adiabatic conditions radial gradients can

    often be ignored, but for nonadiabatic exothermic

    =I.

    r f

    reactions where radial gradients can be large, failure

    The reference time t,,

    was chosen to be the resi-

    to mode l the radial dimension may render the model

    useless. As stated earlier, the reactor in this research

    dence time based upon the initial gas velocity L/v,.

    was designed to minimize heat loss and thus large

    The reference temperature was taken to be the reactor

    radial gradients as done by Be ll and Edgar (1991). A

    inlet temperature

    TO

    These variable definitions led to

    1-D model was therefore developed, which also mini-

    the dimensionless group s given in Table 1.

    mized the number of states upon discretization of the

    The dimensionless boundary conditions were:

    distributed parameter system.

    The simplifications mentioned thus far had the

    greatest impact upon the size of the model a nd were

    implemented primarily for this reason. Other assump-

    tions described below were based solely upon physical

    considerations. Because the reactor was operated at

    low pressures, the process gases were assumed to

    a f

    i =o:

    - =

    Pe,(? - fO ),

    ai

    3Y,

    c - = Pe,(Yco - YO),

    a2

    K?Yco

    i= 1:

    -O_

    ai- di

    (3)

    (4)

    (5)

  • 7/24/2019 Nonlinear Model Predictive Control of A

    6/20

    The zero gradient boundary condition employed at

    the reactor exit for the above equations is a numerical

    approximation to experimentally observed behavior.

    However, this is a comm on assumption even when it

    cannot be experimentally verified. As the reactive

    gases move from the active catalyst bed to the inert

    support, reaction ceases. Thus, the effluent concen-

    tration becomes fixed, and for adiabatic systems so

    does the temperature. The initial conditions were

    similarly nondimensionalized and resulted in two

    dimensionless profiles, one for each equation.

    A num ber of investigators including Moe (1962),

    Ampaya and Rinker (1977), L ee (1980), Bo nvin

    (1980), Newsom e (1980), and Bell and Edgar (1991)

    have studied a n array of shift catalysts. Many rate

    expressions for the reaction h ave been proposed, but

    in this research special consideration was given to

    expressions that account for reaction equilibrium

    effects. The following second order rate expressions

    provided by United Catalysts, Inc. was employed.

    (-rco) = kt&J.Y rilo - JJc,,-YH* I&l(~)l

    This expression has the advantage of directly incor-

    porating the effects of steam. This is especially useful

    if the flow of steam to the reactor is adjustable. The

    rate constant k was assumed to have an Arrhenius

    temperature dependence.

    4.

    SOLUTION TECHNIQUES

    The model equations developed in the previous

    section were solved n umerically using a Galerkin

    finite element technique. A linear combination of

    piecewise-simple polynomials with respect to some

    suitably chosen partition of the spatial domain con-

    stitute a finite-dimensional approximation to the true

    model solution. Th is approximation ultimately led to

    a differential-algebraic equation (DAE) set, which

    was integrated using an implicit, predictor-corrector

    integration scheme. T welve piecewise linear elements

    were used to spatially discretize the WGS reactor

    88

    G. T. WRIGHT and T. F. EDGAR

    model, resulting in 12 differential and 14 algebraic

    states. DAEs are notoriously difficult to initialize for

    integration since finding a consistent set of initial

    conditions is nontrivial. In this research, consistent

    initial conditions were determined in a way that

    permitted both steady-state and dynamic start-up.

    The technique is outlined using an explicit DAE, but

    implicit DAEs may be initialized sim ilarly:

    k = f(x, Y), (6)

    0 = g(x, Y). (7)

    The differential states of the DAE, x were given

    values xt,, determined by some initial (perhaps arbi-

    trary) profile. y0

    was then determined so that

    equation (7) was satisfied. Having determined both

    the differential and algebraic states, the derivatives

    were determined to satisfy equation (6). This, of

    course, permitted dynamic start-up. Equally import-

    ant, the same technique was employed for steady-

    state start-up. How best to choose x0 was addressed

    as follows.

    For the WGS reactor model, the differential states

    corresponded to spatially distributed centerline-bed

    temperatures, which were measurable. Spatially dis-

    tributed, radially averaged bed compositions com-

    prised the algebraic states, but only the composition

    at the reactor exit was measurable. The seven equally-

    spaced temperature m easuremen ts collected axially

    along the reactor w ere used to determine approxi-

    mate model temperatures at the spatial nodes via

    linear interpolation. These w ere used to initialize the

    model for experimental applications.

    The nonlinear algebraic equation sets that arise

    when implementing the above procedures were solved

    using HYBRD, a subroutine from the MINPACK

    libraries (More et al., 1980). Caracotsios DASAC

    (1986) was used to integrate the model in time and to

    evaluate state and output sensitivities with respect

    to the sequence of manipulated variable moves.

    DASAC is based upon the predictor-corrector inte-

    gration algorithm, DASSL, developed by Petzold

    (1983).

    5. PARAMETER FSTIMATION AND MODEL

    VERIFICATION

    For the WGS reactor model, the parameter set of

    to be fitted consisted of the following parameters:

    E,,

    activation energy.

    A, pre-exponential factor.

    PY

    heat capacity of solid medium.

    These parameters were chosen because they were

    poorly known and could not accurrately be deter-

    mined a ~riori. The first two of these parameters are

  • 7/24/2019 Nonlinear Model Predictive Control of A

    7/20

    Nonlinear

    model

    predictive control

    89

    clearly kinetic rate parameters and the last was used

    primarily for fitting reactor dynamics since it appears

    only in the Lewis number which is the coefficient for

    the time derivative of dimensionless temperature. The

    need to estimate heat transfer parameters was re-

    moved by virtue of the adiabatic a ssumption which

    eliminated the reactor wall energy balance. This

    assump tion proved to be quite reasonable since heat

    losses to the surrounding were small relative to the

    heat generated by reaction.

    Seven temperature measurements located axially

    along the centerline of the active bed were used for

    parameter estimation. The m easured effluent CO

    composition was also employed initially, but sub-

    sequent studies proved the parameters to be insensi-

    tive to this value when used with the multiple

    temperature measurements.

    The model states were most sensitive to the rate

    parameters on the interior of the active bed for

    high-temperature operation and at the exit for low-

    temperature operation. The state sensitivities with

    respect to the activation energy and the pre-exponen-

    tial constants varied roughly in proportion to one

    another. Linearly dependent sensitivities lead to vir-

    tually dependent first-order necessary conditions for

    the parameter estimation problem and an ill-posed

    estimation problem . For this reason the centering

    technique described by Bates and Watts (1988), which

    improves the conditioning of the estimation problem ,

    was employed. The Arrhenius expression:

    was rewritten as

    ,

    where

    A=Aexp(- ) .

    The mean temperature T , was chose n to lie within

    the range of observed bed temperatures. The primary

    effect of centering was to reduce the collinear depen-

    dence between the sensitivities. What was not clear,

    however, was how best to choose T , for optimal

    conditioning of the estimation problem. The par-

    ameter estimation problem was solved for several

    values of

    T,,,

    . The value that was ultimately employed

    yielded the smallest 2 - d intervals for the parameter

    estimates as determined by GREG (Caracotsios,

    1986), a parameter estimation package developed by

    Caracotsios.

    A weighted least squares cost function was used as

    the measu re of plant-model misma tch in this re-

    search:

    The cost function is equivalent to the maximum

    likelihood estimator when the measurement errors

    are uncorrelated and normally distributed and their

    variances are constant. Since these assumptions do

    not apply to our data, we are content to interpret the

    results simply as weighted least squares estimates.

    Because each temperature measurement was assumed

    to be similarly accurate, the weights u, were each

    given a value of unity. For mea sureme nts of varying

    qualities, how ever, the weights can be- adjusted to

    reflect measu rement confidence. The weights can also

    be used as scaling factors for measurements of differ-

    ent magnitudes. This was unnecessary here since the

    equations and the data were scaled via nondimen-

    sionalization.

    The parameters were estimated using eight steady-

    state data sets and three dynamic data sets. The

    dynamic experiments consisted of perturbations to

    the active bed inlet temperature usually in the form

    of first-order exponentially filtered steps. Flowrates

    ranged from approx. 9.6 to 12.0 SLM, and the inlet

    conditions varied as indicated in Table 2 for the

    steady-state experiments. Experiments rb1028a,b,

    and c were performed to verify reproducibility o f

    results from

    several months earlier. Parameter

    estimates obtained from steady-state data for the

    activation energy and the pre-exponential constant

    are given in Table 3. A lso listed are the 2 - cr

    intervals associated with each parameter estimate.

    The 2 - D interval is a simple measure of the quality

    of the parameter estimate-small values relative to

    Table 2. Operating conditions for steady-state experiments

    Inlet

    Inlet mol fraction

    tcmperaturc

    Experiment ID C

    co

    H>O CO1

    HZ

    Total

    flowrate

    SLM

    rbO7231a

    rbO723lb

    rbO7241a

    rbO724lb

    rbO724lc

    rb10281a

    rbl028lb

    rb10281c

    285.00 0.154 0.534

    0.156 0.156 9.6 I

    280.00 0.154 0.534

    0.156 0.156 9.6 I

    28 I .24 0.167 0.588

    0.167 0.084 11.95

    285.00 0.167 0.588

    0.167 0.084 11.95

    290.00 0.167 0.588

    0.167 0.084 11.95

    281.24 0.165 0.585

    0.166 0.084 I i.97

    285. I7 0.165 0.585

    0.166 0.084 11.97

    290.00 0.165

    0.585

    0.166

    0.084 Il.97

  • 7/24/2019 Nonlinear Model Predictive Control of A

    8/20

    90

    G. T. WRIGHT and T. F. EDGAR

    Table 3. Optimal parameter estimates from steady-state experiments

    Parameter

    A

    E.

    T, = 297

    SE of residuals

    Estimated 2-u

    value

    Interval

    1.055 x 10-s 4.258 x lO-7

    2.783 x IO+ 1.189 x lo+

    7.199X 10-3

    the estimates are preferred. These intervals are strictly

    valid only if the parameter estimates are independent

    and normally distributed. Figure 2 illustrates the

    good experimental data/model agreeme nt for three of

    the eight steady-state experiments. Th e sam ple given

    represents low; m edium- and high-temperature oper-

    ation.

    As stated earlier the heat capacity of the solid

    phase was estimated to obtain a good dynamic fit.

    The rate parameters were also re-estimated using the

    steady-state parameter estimates as initial guesses.

    Table 4 lists the parameter estima tes obtained w hen

    the dynamic experimental data were employed. Be-

    cause the rate parameters varied only slightly from

    the values obtained using steady-state data and since

    all parameters were well determined, we may con-

    clude that the good dynamic fit illustrated in Fig. 3

    for the seven equally spaced axial centerline bed

    temperature measurements was primarily achieved

    via the solid heat capacity estimate. Note that the

    dynamic data led to sma ller 2 - o intervals for the

    kinetic parameters. Figure 3 shows experiment

    rb7241d. The rem aining dynamic experim ents be-

    haved similarly. We may also conclude that the mode l

    was valid over the nonlinear operating space given by

    the conditions in Table 2.

    6. CONTROLLER DEVELOPMENT

    There are two ways of performing model-predictive

    control calculations. T he first method is sequential

    and employs separate algorithms to solve the differ-

    ential equations, and carry out the optimization.

    First, a manipulated variable profile is guessed, and

    the differential equations are solved numerically to

    obtain an open-loop variable profile. Based upon the

    numerical solution, the objective function is evalu-

    ated. The gradient of the objective function w ith

    respect to the manipulated variable is determined

    either by finite differencing or by solving sensitivity

    equations. Finally, the control profile is updated

    using some optimization algorithm, and the process

    repeated until the optimal profiles are obtained. This

    constitutes a sequential solution and optimization

    strategy, and recent versions of this strategy have

    been reported by: Asselmeyer (1985), Morshedi

    (1986), Jang et al. (1987), Kiparissides and Georgiou

    (1987) and Peterson er al. (1989). The av ailability of

    accurate and efficient integration and optimization

    packages permits implementation o f this method

    with little programm ing effort. However, constraint

    handling is poorer than in an alternative method

    which uses a simultaneous solution and optimization

    strategy.

    When the second or simultaneous approach is

    adopted, the model differential equations are dis-

    cretized,

    and along with the algebraic model

    equations are included as constraints in a nonlinear

    programming (NLP) problem. The optimization

    of the objective function is performed such that

    460

    440

    420

    E

    3

    380 t

    z

    P

    360

    p1

    340

    t

    model - rb7241a -

    model - rb724lb ----.

    model - rb7241c -----

    rb724la

    rb7241b +

    rb7241c D

    320

    I

    I .

    0 0.2 0.4 0.6

    0.8

    1

    Normalized Axial Length

    Fig. 2. Steady -statemodel predictions

    nd experimental observations

    for experiments b724

    a,

    rb724

    b

    rb724 lc.

  • 7/24/2019 Nonlinear Model Predictive Control of A

    9/20

    Nonlinear model predictive

    control 91

    Table 4. Optimal parameter estimates from dynamic experiments

    timization. The num ber of NLP con straints arising

    Estimated 2-a

    Parameter value Interval

    from sequential optim ization and solution is indepen-

    A 1.070 x 10-S 3.210 x IO-

    dent of the prediction horizon.

    6

    2.734 x 1O+4 9.953 x IOf

    Cp* 4.480 x 10-l 6.123 x IO-

    Having selected a sequential optimization and sol-

    7- =297C

    ution strategy, we opted for a feasible path approach

    m

    SE of residuals 7.282 x 1O-3

    vs

    an infeasible path strategy. Infeasible path strat-

    egies do not require that the constraints be satisfied

    (the model equations be solved) at each iteration, but

    the discretized model differential equations are sat-

    find the optimum and satisfy the constraints simul-

    isfied and other constraints on the states and manip- taneously. Feasible path strategies, on the contrary,

    ulated variables are met. Key results em ploying satisfy the constraints at each iteration while seeking

    this method have been reported by Hertzberg and

    the optimum. The sequential optimization and sol-

    Asbjornsen (1977), Biegler (1984), Cuthrell and ution strategy is necessarily a feasible path technique

    Biegler (1987), Renfro et al. (1987), P atwardhan et al. (at least in terms of the model equations), but the

    (1988, 1989, 1990, 1991) and Eaton and Rawlings simultaneous optimization strategy can employ either

    ( 1990). a feasible or infeasible path solution strategy. While

    In this work, a sequential optimization and sol- infeasible path strategies appear to be computation-

    ution strategy was employed. We justified this choice

    ally more efficient, the feasible path techn ique offers

    for experimental application based predominantly

    several advantages. The first of these is that if the

    upon the dimensionality of the NLP which aro se optimizer sh ould fail, the controller need only com-

    for the two strategies. The num ber of constraint pare the values of the objective function at the

    equations arising from discretization in the simul-

    beginning of the optimization to the value at failure.

    taneous optimization and solution strategy varies

    If this value improves, the controller output can be

    directly with the prediction horizon (PH), an integer

    implemented. When infeasible path strategies fail,

    multiple of the sampling time. If the model order is

    there is no direct recourse since the model is usually

    sufficiently large, the computational burden of a large infeasible. Another advantage of feasible path tech-

    prediction horizon becomes more intense than that

    niques is that one enjoys the luxury of intentionally

    associated with the integrations required in the

    solving the NLP suboptim ally. This becomes import-

    sequential strategy. Moreover, the relative increase in

    ant for real-time implementation since the solution of

    the computation time required wh en the prediction

    each NLP must not exceed som e fixed time, usually

    horizon is extended is smaller for the sequential

    the sampling interval. If the time limit is approached

    strategy than for the simultaneous approach to op-

    in the course of solving the NLP, optimization can be

    460

    440

    420

    400

    z

    -

    t

    380

    2

    ::

    360

    H

    d

    340

    320

    300

    280

    Experimental Data -

    Model Data ----.

    0

    50 100 150 200 250 300 350

    400 450

    Time minutes)

    Fig. 3. Dynamic model predictions and experimental observations for experiment rb7241d (see Table 2,

    experiment number 3 for conditions).

  • 7/24/2019 Nonlinear Model Predictive Control of A

    10/20

    92

    G. T. WRIGHTand T. F. EDGAR

    halted and the solution from the last complete iter-

    ation can be implemented.

    6.1. Sequenti&

    so l u t i o n an d op t im i z a t i o n st r a t e gy

    The nonlinear model predictive controller im-

    plemented in this research was formulated as the

    following NLP:

    mip @[x(ri), u(zi)] i = 1,2,. _ , PH,

    subject to satisfying:

    1.

    2.

    3.

    4.

    5.

    6.

    Model differential and algebraic equations:

    E(t)* = f]x(O, uU)l> 8)

    where x(2) E W,

    u(t) E 4t, E(t) E W X ,

    rank[E(t)]

  • 7/24/2019 Nonlinear Model Predictive Control of A

    11/20

    Nonlinear model

    sequence of piecewise constant inputs are denoted by

    x,,. In order to determine the gradient of the objective

    function with respect to v, the sensitivity equ ations

    for the states of the DAE must be determined.

    Let W represent the sensitivity matrix of the state

    vector with respect to u,:

    w, = x*,.

    Then the dynamic evolution of W, is determined by:

    E(t)W, = J(x, v)W, + B,(x, v),

    i = 1, . . . , CH,

    (11)

    where

    Bj(Xv VII =

    f (x. q 1,

    zj-, < t -c t,,

    0

    3

    otherwise,

    and J(x, v) is the Jacobian of f(x, v). The sensitivity

    equations are subject to the initial conditions:

    W,(r,) = 0.

    7. EXPERIMENTAL CONTRO L STUDIES

    As a consequenc e of the computational burden

    associated with NMP C, a slow sampling rate was

    required to accomm odate the relatively long compu-

    tation times evolving from the solution process.

    Because this factor also diminished the disturbance

    rejection capabilities of NMPC , NMPC was im-

    plemented in a master-slave cascade control

    configuration where a low-level linear controller was

    a significantly faster sampling rate was employed.

    As described previously, the primary input to the

    reactor was power to the inlet feed heater. W hile

    varying the power level affected all bed temperatures

    and compositions, we focused specifically upon its

    impact on the active bed inlet temperature. Power

    was never determined directly as the NMP C output.

    Instead, NMPC determined a target value for the

    active bed inlet temperature that would presumably

    lead to the desired bed behavior. Recall that the

    active bed inlet temperature constituted the inlet

    boundary condition for the reactor energy balance.

    This relationship was used to construct the NMPC

    control strategy for the WGS reactor illustrated in

    Fig. 4.

    7.1. The WGS r eac t o r i n l e t t empe r a t u r e f oop

    The open-loop reactor inlet temperature behavior

    was influenced predominantly be feed-gas flowrate.

    The inlet behavior was quite linear for flowrates

    ranging from 9 to 13 SLM. Time constan ts varied

    from 10.3 to 14.3 min and the dead time was approx.

    3 min. The static gain w as 2.8 with a maximum

    variation of 15%. Since the reactor inlet section w as

    predictive control

    __......................

    93

    Feed

    R

    e

    a

    :

    0

    a

    -r ----------

    *

    Fig. 4. Cascade control con figuration for implementing

    model-based control strategies.

    packed with inert Pyrex glass beads, no reaction

    occurred in this region. Therefore, heating and cool-

    ing of the reactor inlet was virtually independent of

    feed composition. However, when the reactor feed

    stream was composed of a reactive gas mixture (e.g.

    the gas mixture contained CO), a portion of the heat

    generated from reaction in the active bed diffused

    upstream to the inlet, marginally impacting reactor

    inlet behavior. Because the static gain, time constant

    and time delay varied little for the flowrates of

    interest, PID control was used to close the loop.

    In light of the master-slave, cascade control

    configuration, it was imperative that the PID con-

    troller (the slave) effectively track the reactor inlet

    temperature target value computed by a model-based

    controller (the master). The PID controller was

    tuned, therefore, using an ITAE tuning rule for

    set-point tracking. Figure 5 is typical of the closed-

    loop response, which h ad the appearance of a first-

    order plus dead-time transfer function step response.

    The closed-loop time constant for the flowrates of

    interest was approx. 6 min and the dead-time was

    3 min.

    We have already concluded that given flowrate an d

    composition, the relationship between inlet bed tem-

    perature and subsequen t bed behavior was well

    defined. The relationship governing bed behavior as

    a function of power to the inlet heater was poorly

    defined and subject to disturbances which we re not

    easily measured or modeled. Fortunately, the need to

    model the reactor inlet was eliminated by assumin g

    that the PID controller consistently generated a

    first-order plus dead-time closed-loop response as

    described above.

  • 7/24/2019 Nonlinear Model Predictive Control of A

    12/20

    288

    287

    286

    285

    284

    48

    46

    G

    T. W RIGHT and T. F. EDGAR

    1

    44

    1

    0 20

    40 60

    80 100

    T i m e m i n u te s )

    Fig. 5. C lose-loop response of the reactor inlet temperature for a 5C set-point increase and a Aowrate

    of 13 SLM.

    7.2 . Closed- loop NM PC exper im ents

    For all NMPC experiments the control horizon

    CH was unity, permitting only one manipulated

    variable move over the entire time horizon. The

    prediction horizon PH was 24 sampling intervals or

    120 min. Aggressive control is generally achieved for

    large CH and small PH. Our objective, however, was

    not to demonstrate aggressive control, but smooth

    consistent control over a broad operation region.

    Therefore, the tuning parameters were selected ap-

    propriately. Furtherm ore, a larger control horizon

    would have made the problem computationally in-

    feasible for real-time application on the WGS system

    because it would have been accompanied by an

    increased number of optimization variables.

    Although every attempt was made to make the

    NMPC algorithm computationally efficient, the sol-

    ution of each NLP required from 3 to 4.5 min. In

    light of this, the sampling interval

    T ,

    for control of

    the sixth bed temperature was chosen to be 5 min.

    This value complied with established guidelines pre-

    sented by Seborg et a l . (1989) and Astrom and

    Wittenmark (1990), who suggest that the sampling

  • 7/24/2019 Nonlinear Model Predictive Control of A

    13/20

    Nonlinear model predictivecontrol

    95

    rate be less than a tenth of the dominant time

    constant or that the ratio of the sampling rate to the

    time constant lie between 0.1 an d OS. The dominant

    time constant for the system was approx. 55 min and

    the

    dead time, 30 min. A nominal dry-gas inlet com-

    position of 40% CO, 40% CO2 and 20% H2 was used

    for all experiments unless otherwise stated. The volu-

    metric dry-gas to steam ratio was 0.625, and the total

    gas flow was 13 SLM.

    The first experiment w as intended simply to

    demon strate that NMP C handles set-point tracking

    smoothly and efficiently. Figure 6 depicts the closed-

    loop response of bed temperature 6 to a 16.5%

    step-change in its set-point. For clarity, we reiterate

    that the manipulated variable for the NMP C loop

    was the inlet temperature set-point, and that the

    manipulated variable for the PID

    loop was power to

    the inlet heater. NMP C was permitted to change the

    325

    I

    I

    320

    G

    .m

    t

    ~______________________________._____~

    300

    240

    288

    286

    284

    282

    280

    278

    276

    3

    44.0

    2

    H

    42.0

    set

    -Point -----

    Outpuf -

    ____________I

    4

    1

    set-POi

    nt

    output

    I

    1

    c.

    I

    0 50

    100 15

    200 250

    300

    Time (minutes)

    Fig. 6. NMPC experimentnumbe r I: WGS reactor response o a 16.5C step increase n the set-point

    for bed temperaturenumber 6.

  • 7/24/2019 Nonlinear Model Predictive Control of A

    14/20

    96

    G. T. WRIGHT and T. F. EDGAR

    inlet temperature set-point by no more than +2.5 C

    per control interval. Absolute limits of 270 and 300C

    were also enforced.

    Although small oscillations of f 1C persisted, it is

    apparent that NM PC achieved the desired set-point.

    These minor oscillations were the direct result of

    small inlet temperature oscillations about the inlet

    temperature target value. Since NMP C is model-

    based, dead-time compen sation is inherent, provided

    360

    350

    340

    330

    320

    300

    290

    286

    284

    42.0

    36.0

    the model accounts for it. For the WGS reactor, a

    temperature variation at the inlet initiates a thermal

    wave, which amplifies as it propagates through the

    active bed. This phenome non effectively creates a lag

    that would be modeled as pure delay in a transfer

    function representation of the system. Notice that

    NMP C required only three sampling intervals to

    determine the inlet temperature that would drive bed

    temperature 6 to set-point. Furthermore, once this

    I

    1

    1 1

    Set Point

    ourput -

    Set Point

    output -

    0 100

    200 300

    400 500

    Time (minutes)

    Fig. 7. NMPC experiment number 2: WGS reactor response to a sequence of set-point increases for bed

    temperature number 6 which spans the operating space for nominal feed composition and flowrate.

  • 7/24/2019 Nonlinear Model Predictive Control of A

    15/20

    Nonlinear model predictive ontrol

    value had been determined, manipulated variable

    changes virtually ceased, despite the initial absence of

    response of temperature 6. This example and others

    that follow powerfully illustrate the inherent dead-

    time compen sation of NMP C.

    The second experiment was intended to illustrate

    the ease with which NMPC progresses from a

    state of virtually no reaction to a state of almost

    complete reaction when applied to the WGS reactor.

    This example highlights the effective use of NMPC

    for plant start-up. Figure 7 shows a sequence of

    set-point changes. The first was an 18S C step-

    increase from 306.5 to 325C and the second a

    25C step-increase to 350C. A velocity constraint

    permitted NM PC to manipulate the inlet temperature

    set-point by n o m ore than + l.O C per control

    interval.

    We note at this point that when Ziegler-Nichols

    tuning rules are adopted for PID tuning, positive

    static gain variations should not exceed approx.

    lOO%, and even this value is borderline. While more

    advanced PID tuning strategies have been de-

    veloped more recently, this rule of thumb still loosely

    applies.

    As with the previous example delay time was easily

    accomm odated as evidenced by the absence of exces-

    sive manipulated variable move ment. More signifi-

    cant, however, was the successful handling of the

    static gain variations. Figure 7 clearly illustrates

    that for an 18.5C change in the output, an inlet

    temperature change of approx. 5C was required.

    For the subsequent 25C output increase, which

    occurred at higher CO conversion, an inlet tempera-

    ture change of approx. 2.4C was required, a tripling

    of the static gain. NMPC inherently recognized these

    gain varitions and responded accordingly. For the

    same operating conditions, Fig. 8 illustrates the

    poor simulated response achieved using a PID con-

    troller, tuned with ITAE rules for set-point tracking.

    The third and final NMP C experiment dealt

    with the disturbance rejection capabilities of NMPC.

    First, steady-state was achieved with an output set-

    point of 310C. At time equal to 30 min, the dry-gas

    flowrate was decreased by 10% to 4.5 SLM. The

    dry-gas com position and steam flowrate were not

    altered. Since flowrate and composition measure-

    ments appeared as parameters in the NMP C model,

    an inherent feedforward action caused an immed iate

    drop in the inlet temperature set-point (Fig. 9, arrow s

    mark the flowrate decrease). This occurred before any

    significant response in bed temperature 6, the feed-

    back variable. In fact, the output only began to

    respond approx. 10 min later.

    This disturbance rejection example highlights

    a flaw of the NMP C control implementation.

    When constructing the controller, it was assumed

    that the inlet temperature loop had a perfectly

    consistent first-order response for set-point track-

    ing. Figure 9 clearly demonstrates that this

    assumption was violated in the presence of a dis-

    turbance. In the next section we discuss the rami-

    fications of this assump tion by examining the

    model states with and without inlet temperature

    feedback.

    97

    370

    360

    350

    iz

    z 340

    2

    ::

    1 330

    300 I

    1

    0 100

    200 300 400 500 600

    Time (minutes)

    Fig. 8. Simulated WGS reactor response to a sequence of set-point increases for bed temperature

    numb er 6, wh ich spans the operating space for nominal feed composition and ilowrate.

  • 7/24/2019 Nonlinear Model Predictive Control of A

    16/20

    98

    G T. WRIGHTZUI~ T F EDGAR

    set-Polnc ----

    output - _

    320

    318

    316

    314

    312

    310

    308

    306

    286 0

    265 0

    284 0

    282 0

    281 0

    260 0

    40 0

    39 5

    39 0

    36 5

    36 0

    37 5

    37 0

    36 5

    36 0

    35 5

    35 0

    Set-Point ----.

    output -

    1

    1 1

    0

    so

    100 150 200

    Time lminucesl

    Fig. 9. NMF C experiment number 3: WGS reactor response to a 10% step decrease in the nominal dry-gas

    flowrate.

    7 . 3 . Compa r i son o f p l an t and mode l ou t pu t s f o r

    NMZC

    Experiment three clearly demonstrated a violation

    of the assumption that the closed-loop behavior

    of the inlet temperature loop was first-order. The

    unexpected response of bed temperature 1 was a

    consequence of the 10 step-decrease in the dry-gas

    flowrate. Figure 10 compares the output response

    (bed temp erature 6) actually experienced in the plant

    to the model response. Notice that the model tem-

    perature increased slightly d ue to the decreased dry-

    gas flowrate (arrows mark the flowrate decrease).

    When compared to the actual temperature increase in

    the plant, however, the model temperature increase

    was small.

    The temperature increase in the plant output was

    the cumulative effect of increased residence time,

  • 7/24/2019 Nonlinear Model Predictive Control of A

    17/20

    330

    325

    310

    Nonlinear model predictive control

    I

    Plant

    -

    Model vf Inlet Temp. Feedback

    __-_.

    Model w/o Inlet Temp. Feedback --.--

    99

    0

    50 100

    150 200

    Time (minutes)

    Fig. 10. Comparison of plant and model states with and without inlet temperature feedback when a 10

    step decrease in the nominal dry-gas flowrate is applied to the reactor system.

    which permitted more reaction, and increased inlet

    temperature caused simply be a slower rate of heat

    removal. Only the first of these effects w as considered

    in the model. Incorportion of the second effect would

    have required that the reactor inlet section be mod-

    eled. Such a model would only increase the overall

    reactor model size while adding inform ation that can

    be otherwise accounted for. Figure 10 shows a closed-

    loop simulation where the actual reactor inlet tem-

    perature is used as an input to the model. In this case,

    the model-output time derivative was almost pre-

    cisely that experienced by the plant. A sm all, slowly

    varying bias ranging from 5 to 10C persisted, but the

    feedback mechan ism of NMPC is designed to effec-

    tively deal with this phenom enon. We conclude,

    therefore, that better plant-model agreement is

    achieved if the actual inlet temperature is used as an

    input to the model. Inlet temperature feedback to

    reset the model boundary conditions would effec-

    tively achieve a feedforward control strategy (Wright,

    1992).

    Figures 11 and 12 illustrate that an inlet tempera-

    ture feedback would be much less significant for

    set-point tracking. In fact, for experiment two there

    was no substantial distinction between the model

    output with or without feedback. As with experiment

    three, the bias varied slowly here, increasing with high

    temperature operation. In experiment one, use of

    380

    1

    1

    __----_

    --._

    370 -

    ,,~~:___---~-~ ---r..__

    ../-

    -9

    360 -

    ,.;?

    :,s

    :

    u

    350 -

    E

    J

    E

    340 -

    P

    8

    330 -

    Plant

    -

    Model w/ Inlet Temp. Feedback

    _____

    Model w/o Inlet Temp. Feedback ----_

    l-

    300 1 i

    0 100

    200 300

    400 500

    Time (minutes)

    Fig. 11. Comparison of plant and model states with and without inlet temperature feedback when a

    sequence of set-point changes was applied to bed temperature 6.

  • 7/24/2019 Nonlinear Model Predictive Control of A

    18/20

    G. T. WRIGHT and T. F.

    EDGAR

    325

    Model w/

    Inlet Temp. Feedback

    __--.

    Model w/o Inlet Temp. Feedback -----

    300 1 I 1 1

    0 50 100 150 200 250 300

    Time (minutes)

    Fig. 12. Com parison of plant and mode l states with and without inlet temperature eedback when a

    16SC step-point ncreasewas appliedto bed tem perature .

    inlet temperature as a mo del input had a marginally

    greater effect than for experiment two. Even so, the

    output behavior in the plant could not be captured

    for the time interval ranging from 100 to 175 min.

    This example illustrates, however, that NMP C is

    robust to plant-model mismatch. It is clear, that in

    the absence of repeated disturbances, both techniques

    lead to the same model output, but inlet temperature

    feedback may significantly affect transient behavior.

    7.4.

    Compar i son w i th c l osed - l oop GPC

    Adaptive GPC was implemented using the control

    structure outlined for NMP C. Unlike the NMP C

    experiments, how ever bed tempera ture 4 w as taken to

    be the controlled variable. Because the computation

    time required for adaptive GPC was small, a

    sampling time of 2 min, based solely upon the open-

    loop dynamics of bed temperature 4, was adopted.

    For an inlet temperture of 280C and nominal values

    of composition and flowrate, the dead-time was

    approx. 14min, the time constant 25 mm , and the

    gain 1.5. The gain increased by approx. 120% from

    this low reaction state to a state of com plete reaction.

    The control experiment described below used a

    prediction horizon of 20 sampling intervals, and a

    control horizon of unity. A move supression factor of

    10 was also employed, and GPC was permitted to

    change the set-point by no more than _t 1C per

    control interval. The recursive least squares esti-

    mation algorithm of Chen and Norton (1987) was

    employed for parameter estimation. The model took

    the form:

    A (4 -ly(t) = q -B(q -)24(t) + ci,

    (12)

    where A (q -) and B(q - ) are polynomials in the

    backward shift operator of orders 1 and 3, respect-

    ively.

    Figure 13 illustrates a sequence of three 5C step-

    increases in the target value for bed temperature 4.

    While the close-loop response for the first and second

    increments were satisfactory, it is clear that the

    response became progressively worse with increasing

    operating temperature. The oscillatory behavior was

    obtained despite parameter adaptation. In addition,

    this control pro blem was less challenging than the

    problem to which NMP C was applied. We con-

    cluded, therefore, that traditional adaptive control is

    not well suited for WG S reactor start-up, since the

    linear model, even with parameter adaptation, does

    not a dequately reflect the rapidly changing nonlinear

    dynamics of the system.

    8.

    CONCLUSlONS

    The primary goal of this research was to develop

    an advanced nonlinear control strategy for fixed-bed

    catalytic reactors. The control method was applied

    experimentally using a laboratory-scale water-gas

    shift WGS reactor. The following conclusions may be

    drawn from the results of this work.

    An adiabatic, pseudo-hom ogeneous WGS reactor

    mode l represented the physical system well over the

    operating space of interest. The physically reason-

    able, simplifying assusmptions that were adopted

    proved useful in developing a low-order m odel, suit-

    able for implementation in an NM PC framework.

    The Galkerin technique on finite elements with piece-

    wise linear polynomial approximations proved not to

  • 7/24/2019 Nonlinear Model Predictive Control of A

    19/20

    Nonlinear model predictive control

    306

    302

    296

    286

    284

    Set Point

    output -

    46.0

    3 50 100 150 200 250 300 350 400

    450

    Time (minufcr)

    Fig. 13. Adaptive GFC experiment: W GS reactor response to a sequence of set-point increases for bed

    temperature numbe r 4 for nominal composition and flowrate.

    be susceptible to oscillatory behavior over the spatial

    domain when 12 nodes were employed for discretiza-

    tion. Estimation of dynamic and steady-state

    parameters were efkctively decoupled for the pseudo-

    homo geneous W GS reactor model. Furthermore, in-

    formation-rich dyn amic data yielded good parameter

    estimates w ith less experimental effort.

    The control experiments demonstrated that absol-

    ute plant-model agreeme nt was not imperative for

    good control using NMPC . However, temporal first-

    derivative information,

    consistent with

    plant

    behavior, wa s crucial to good performance. NMP C

    was better suited for feedforward dynam ic co mpen-

    sation than linear techniques since the nonlinear

    model has an inherent characterization of the feed-

    forward mechanism. Feedforward control signifi-

    cantly

    enhances NMP C performance. Finally,

    NMP C was effectively used to start-up the WGS

  • 7/24/2019 Nonlinear Model Predictive Control of A

    20/20

    102

    G. T. WRIGHT and T. F. EDGAR

    system.

    NMPC

    was

    superior in this regard to tra

    ditional control techniques since broad nonlinear

    operating regions were traversed. Adaptive linear

    control appears to be unsuitable, since parameter

    estimates varied a s rapidly a s the state, making

    successful parameter adjustment extremely difficult.

    REFERENCES

    Ampaya J. P. and R. G. R inker, Autothermal reactors with

    internal heat exchange. J. C h e m . E n g n g S c i . 3 2 , 13 2 7

    ( 1 9 7 7 ) .

    Asselmeyer B., Optimal control for nonlinear systems calcu-

    lated with small computers.

    J . O p t . T h e or y A p p l i c . 4 5 , 5 3 3

    ( 1 9 8 5 ) .

    Astrom K. J. and B. Wittenmark, C o m p u t e r C o n t r o l l e d

    S v s t em s : T h e or v a n d D e s n . 2nd Edn. Prentice-Hall. New

    Jkrsey ( 1 990 ) . _

    Bates D. M. and D. G. Watts, N o n l i n e a r R e g r es si o n A n a l y si s

    a n d I t s A i c a t i o n s . Wiley, New York (1988).

    Bell N. H.-&d T. F. Edgar, Modeling. of a fixed-bed

    water-gas shift reactor 1: Steady-state model verification.

    J. P r o c e ss C o n f r o l 1 , 2 2 - 3 1 ( 1 9 9 1 ) .

    Biegler L., Solution of dynamic optimization problems by

    successive quadratic programming and orthogonal collo-

    cation.

    C o m p u t e r s

    them. E ngng 8, 243 (1984).

    Bonvin D., Dynam ic modeling and control structures for a

    tubular autothermal reactor at an unstable state. Ph.D.

    Thesis, University of California at Santa Barbara (1980).

    Brockett R. W ., Feedback invariants for nonlinear systems.

    In P r o c . 7 t h Z F A C W o r l d C o n g r . , Helsinki (1978).

    Caracotsios M., Model parametric sensitivity analysis and

    nonlinear parameter estimation-theory and appli-

    cations. Ph.D. Thesis, The University of Wisconsin at

    Madison (1986).

    Newsom e D. S., The water-gas sh ift reaction. Cu f a l . R e v . -

    S c i . Engng 21, 275 (1980).

    Patwardhan A. A. and T. F. Edgar, Nonlinear model-

    predictive control of a packed distillation column. In A m .

    C o n t r o l C o n f . , Boston, MA (1991).

    Patwardhan A. A ., J. B. Rawlings and T. F. E dgar, Model

    predictive control of nonlinear processes in the presence

    of constraints. In

    A Z C h E J I A n n u a l M e et i n g ,

    Washington,

    DC (1988).

    Patwardhan, A. A., J. B. Rawlings and T. F . Edgar, Model

    predictive control of nonlinear processes in the presence

    of constraints. In Z F A C J Z E E E S ym p . o n N o n l i n e a r C o n -

    t r o l S ys t em s D e s i g n , pp. 45&460 (1989).

    Patwardhan A. A., J. B. Rawlings and T. F. Edgar,

    N o n l i n e a r m o d e l predictive control. C h e m . E n g n g C o m -

    m u n . 8 7 , 1 2 3 ( 1 9 9 0 ) .

    C a r b e r r y

    J. C.

    a n d M . M .

    Wendel, A

    c o m p u t e r

    model of the

    fixed bed catalytic reactor: the adiabatic and quasi-adia-

    batic cases. A Z C h E J f 9 , 1 2 9 - 1 33 ( 1 9 6 3 ) .

    Chen M. J. and J. P. Norton, Estimation technique for

    tracking rapid parameter changes. ht. J . C o m r o l 45,

    1387-1398 (1987).

    Clarke, D. W., C. Mohtadi and P. S. Tuffs, Generalized

    predictive control-part i the basic algorithm. A u t o m a t i c

    2 3 , 1 3 7 - 1 4 8 ( 1 9 8 7 ) .

    Cuthrell J. E. and L . T. Biegler, On the optimization of

    different/algebraic process systems. A Z C h E J I 3 3 , 1 2 57

    ( 1 9 8 7 ) .

    Cutler C. R. and B. L. Ram aker, D ynamic matrix control-

    a computer control algorithm. In J o i n r A u t o m a t i c C o n t r o l

    C o n f . P r o c . P a p e r W P - 5 B San Francisco, CA (1980).

    Danckwerts P. V., Continuous flow systems-distribution

    of residence times. Chem. Engng Sci. 2, 1-13 (1953).

    Eaton J. W. and J. B. Rawlings, Feedback control of

    chemical processes using on-line optimization techniques.

    C o m p u t e r s t h e m . E n g n g 1 4, 169-479 (1990).

    Eaton J. W. and J. B. Rawlings, Model predictive control

    of chemical processes. Chem. Engng S c i . (1991).

    Hertzberg T. and 0. A. Asbjomsen , Parameter estimation

    in nonlinear differential equations. In

    C o m p u t e r A p p i i -

    c a t i o n s i n t h e A n a l y s i s o D a t a a n d P l a n t s . Science Press,

    Princeton (1977).

    Peterson T. E., E. Hern andez, Y. Arkun and F. J. Schork,

    N o n l i n e a r

    predictive

    c o n t r o l

    of a semi-batch polymeriz-

    ation reactor by an extended dmc, Paper No. tp4. In A m .

    C o n f r o l . C o n f . , Pittsburg, PA (1989).

    Petzold L. R., A description of DASSL: a differential-

    algebraic system solver. In S c i en t i f i c C o m p u t i n g (R. S.

    Stepleman, Ed.), pp. 65 -68 . North-Holland, Amsterdam.

    Rase H. F.,

    C h e m i c a l R e a c t or D e si g n

    or

    P r o c e s s P l a n t s ,

    Vol. 1. Wiley, N ew York (1977).

    Rawlings J. B. and K. R. Muske, The stability of con-

    strained receding horizon control. Submitted to I E E E

    T r a n s . A u t o . Contr., Sept. (1991).

    Renfro J. G., A. M. Morshedi an d 0. A. Asbjornsen.

    Simultaneous optim ization an d solution of systems de-

    scribed by differential/algebraic equations. C o m p u t e r s

    t h e m . E n g n g 11, SO3 1987).

    Seborg D. E., T. F. Edgar and D. A. Mellichamp, P r o c e s s

    D y n a m i c s a n d C o n t r o l . Wiley, New York (1989).

    U n i t e d C a t a l y s t s C - I 2 O p e r a t i n g I n s t r u c t i o n s .

    Windes, L. C., Modeling and control of a packed bed

    reactor. Ph.D. Thesis, The University of Wisconsin at

    Madison (1986).

    Windes, L. C., M. J. Schwedock and W. H. Ray. Steady

    state and dynamic modeling of a packed bed reactor for

    the partial oxidation of m ethanol to formaldehyde-I,

    model development. C h e m . E n g n g C o m m u n . 7 8 , 1 43 .

    ( 1 9 8 9 ) .

    Hunt L. R., R. Su and G. Meyer, Global transformations Wright G. T., Modeling and nonlinear predictive control of

    of nonlinear systems. I E E E T r a n s . A u t o m a t . C o n t r . A C -

    a fixed-bed water-gas shift reactor. Ph.D. Thesis, The

    2 8 , 2 4 ( 1 9 8 3 ) . University of Texas at Austin (1992).

    Jang S., B. Joseph and H. Mukai, On-line optim ization of

    constrained multivariable processes. A Z C h E J f 33, 26

    (1987).

    Kiparissides C. and A. Georgiou, Finite-element solution of

    nonlinear optimal control problems with a quadratic

    performance index. C o m p u t e r s t h e m . E n g n g 11,77 ( 1 9 8 7 ) .

    Lasdon L. S. and A. D. Waren, G R G U s er s G u i d e . Techni-

    cal Report, University of Texas at Austin (1986).

    Lee A. L., Effect of raw product gas composition on shift

    catalysts. Technical Report Project 61015, DOE Contact

    DE-AC 2I-78ET11330, Institute of Gas Technology.

    Chicago (1980).

    Moe J. M., Design of water-gas shift reactors. Chem. E n g n g .

    P r o c . 5 8 , 33 (1962).

    Mor& J. J., B. S. Garbow and K. W. Hillstrom, User G u i d e

    f o r M i n p a c k - 1.

    Technical Report ANL-80-74, Argonne

    National Laboratory (1980).

    Morshedi A. M ., Universal dynamic matrix control. In

    C h e m i c a l C o n t r o l C o n f . I Z I , S es si o n V I , P a p e r N o 2,

    Asilomar, CA (1986).