Vous êtes sur la page 1sur 17

Awomatifo, Vol. 32, No. 9, pp 128S1301, I!?

36
copylisht @ 1996 l%cviersdencc Ltd
Printed in Great Britain. All rights reserved
ooowo9w6
s1s.wo.oo
PIk SOOOS-1098(96)00086-6

Nonlinear Model Predictive Control of a Simulated


Multivariable Polymerization Reactor Using Second-order
Volterra Models
BRYON R. MANER,+ FRANCIS J. DOYLE III,+ BABATUNDE A. OGUNNAIKE *
and RONALD K. PEARSON *

An MPC algorithm based on the second-order Volterra model is proposed,


anda methodfor obtaining the modeIparametersfromfundumentalmodels
is outlined. Simulation results for two case studies arepresented.
Key Words-Nonlinear control systems; process control; process models; predictive control; inverse
systems.

Abstract-Two formulations of a nonlinear model predictive manipulated and controlled variables. Most MPC
control scheme based on the second-order Volterra series model applications are based on linear finite convolution
are presented. The lirst formulation determines the control ac- models, such as the step and impulse response mod-
tion using successive substitution, and the second method di-
els. An attractive feature of these models is the rel-
rectly solves a fourth-order nonlinear programming problem
on-line. One case study is presented for the SISO control of an ative ease with which they can be obtained from
isothermal reactor which utilizes the fist controller formula- plant data. Hence, these models may be obtained
tion. A second case study is presented for the multivariable con- in a relatively short time without detailed process
trol of a large reactor, and uses the nonlinear programming for- knowledge.
mulation for the controller. The model coefficients for both ex-
amples are obtained by discretizing the bilinear Taylor series ap-
Several researchers have proposed predictive
proximation of the fundamental model and calculating Markov control schemes based on the direct use of nonlin-
parameters. The relationships between discrete and continuous- ear models. In nonlinear quadratic dynamic matrix
time bilinear model matrices using an explicit fourth-order control (Garcia, 1984) a nonlinear model is used
Runge-Kutta method are also included. The responses to set- to compute the effect of past manipulated variables
point changes of both reactors controlled with a linear model
predictive control scheme and the second-order Volterra model
on the predicted output. A linear model, obtained
predictive control scheme are compared to desired, linear ref- by linearizing the nonlinear model at each sampling
erence trajectories. In the majority of the cases examined, the time, is used to compute the manipulated variable
responses obtained by the Volterra controller followed the ref- values. An advantage of this approach is that only
erence trajectories more closely. Practical issues, including the one quadratic program is solved at each sampling
reduction of the number of model parameters, are addressed
in both case studies. Copyright 01996 Elsevier Science Ltd.
interval versus solving a higher-order nonlinear
program. Gattu and Zafiriou (1992) extended this
algorithm by incorporating state estimation using
1. INTRODUCTION a steady-state Kalman filter gain computed at each
sampling time. The performance of the algorithm
Model predictive control (MPC) involves the cal-
was illustrated in the control of a two-by-two con-
culation of a set of manipulated variable moves
tinuous stirred tank reactor (CSTR) and a single-
to minimize an open-loop performance objective
input-single-output open-loop unstable reactor.
over a prediction horizon subject to constraints on
Lee and Ricker (1993) modified Garcia’s algorithm
by implementing an extended Kalman filter and
Received 15 August 1995; revised 20 March 1996; received by incorporating load disturbances on the states as
in final form 5 May 1996. This paper was not presented at any additive stochastic signals. In Gattu and Zafiriou
IFAC meeting. This paper was recommended in revised form (1992), the Kalman filter was designed with white
by Associate Editor B. Wayne Bequette under the direction of
Editor Yaman Arkun. Corresponding author Professor Frances
state noise characterized by a scalar-times-identity
J. Doyle III. Tel. +1 317 494 9472; Fax +1 317 494 0805; E- covariance matrix, which can lead to significant
mail fdoyle@ecn.purdue.edu. bias in the state estimates. Ricker and Lee imple-
t School of Chemical Engineering, Purdue University, West mented their algorithm in an eight-by-eight non-
Lafayette, IN 47907-1283, U.S.A. linear MPC scheme for the Tennessee Eastman test
* E.I. DuPont de Nemours & Co., Wilmington, DE 19880-
0101, U.S.A.
problem (Ricker and Lee, 1995).

1285
1286 B. R. Maner et al.

Li and Biegler (1988) extended the nonlinear in- nonlinear programs that arise in the nonlinear
ternal model control design procedure (Economou internal model control algorithm and nonlinear
and Morari, 1985) to a single-step method that in- MPC based on polynomial ARMA models noted
cluded constraints on the states and inputs. They above. In particular, one group of researchers
demonstrated the performance of their controller (Chow et al., 1994) has developed an algorithm for
for the control of two stirred tank reactors with unconstrained optimization using tensor methods
two inputs and two outputs. The stability of the that employs a fourth-order model of the objective
constrained controller was guaranteed by the small function. In their case studies, the tensor method
gain theorem (Zames, 1966) provided that the pro- required significantly fewer iterations and function
cess was open-loop stable. They extended their al- evaluations to solve most unconstrained optimiza-
gorithm to a multi-step strategy that linearizes a tion problems than standard methods based on
nonlinear model around a nominal trajectory (Li quadratic models. A third advantage is that stabil-
and Biegler, 1989). The modified algorithm was ity results are easier to obtain for Volterra model-
demonstrated with the optimal open-loop control based controllers, since the model depends only
of an SISO and an MIMO (3 x 3) CSTR. In the case on past inputs versus the dependence on both past
studies, it was assumed that there was no plant- inputs and outputs as in the case for polynomial
model mismatch. ARMA model-based controllers. Stability results
Both the nonlinear quadratic dynamic matrix for second-order Volterra model-based controllers
control (Garcia, 1984; Gattu and Zafiriou, 1992; have been presented by several researchers (Genceli
Lee and Ricker, 1993) and modified nonlinear and Nikolaou, 1995; Zheng and Zafiriou, 1993).
internal model control (Li and Biegler, 1988; Li
and Biegler, 1989) algorithms require a fundamen-
tal model of the process. The derivation of these 2. CONTROLLER FORMULATIONS
models can be very time consuming, and can be 2.1. Successive substitution
elusive if the process is not well understood. One The triangular form of the second-order Volterra
nonlinear MPC approach that does not require the model with truncation order N is given by the fol-
availability of a fundamental model uses polyno- lowing expression:
mial ARMA models obtained from input-output
data (Hernandez, 1992). The main advantage of N

using polynomial models is that the one step ahead Y(k) =.Yo + 1 a&k - i)
prediction problem can be formulated as a linear i=l

regression, greatly simplifying the estimation of the


+f’f’bi,iu(k-i)u(k-j). (1)
parameters from input-output data.
i=l j=i
This paper presents a natural extension of the
linear MPC algorithm described in (Ricker, 1985) The triangular form may be used without loss of
to a control methodology based on the second- generality because the second-order parameters are
order Volterra series model. The control action can symmetric for the Volterra model. The yo term is
be computed via successive substitution or through a bias term that appears in the general form of
the direct solution of a fourth-order nonlinear pro- the second-order Volterra model. The ai terms are
gram. A second-order Volterra model can describe impulse response coefficients, and the bi.j terms are
nonlinear behavior such as asymmetric output the second-order Volterra model parameters. The
changes in response to symmetric changes in the first and second terms on the right-hand side of
input. A controller based on this model can yield equation (1) are first and second-order terms which
improved performance over a linear model-based will be denoted as pi[u] and &[u], respectively.
controller. In addition, the model parameters can The block diagram for the second-order Volterra
be obtained from both Carleman linearization of model predictive controller is shown in Fig. 1. P
a nonlinear fundamental model (Rugh, 198 1), and is the actual nonlinear plant being controlled. p:
input-output data (Pearson et al., 1996). Identifi- is an inverse or pseudo-inverse of the linear model
cation of these parameters from input-output data and is identical to the inverse used in linear MPC; 5
enables implementation of this nonlinear MPC is the difference between a vector of desired future
scheme in the absence of a fundamental model. setpoints and a vector of future predicted outputs
Another advantage of a second-order Volterra due to past inputs. It can be seen from Fig. 1 that
model-based controller is that it can be posed as a the second-order Volterra model-based algorithm
fourth-order nonlinear program. While this nonlin- is an extension of the constrained internal model
ear program is more computationally demanding control algorithm described by Ricker (1985). The
than the quadratic program that arises in linear second-order Volterra model replaces the impulse
MPC, it is more readily solvable than the general response model and is used for output prediction.
Voltena MPC of polymerization reactors 1287

Fig. 1. Block diagram for the second-order Volterra model predictive controller.

The corresponding Volterra controller, enclosed 2). . . ., etc. The past-past cross terms of the input
by the dashed box in Fig. 1, replaces the linear appear as g(k + l), g(k + 2)‘. . ., etc. d(k + 1) is the
model-based controller and performs the control difference between the measured value of the pro-
action computations. For the constrained case, a cess output and predicted model output at time k.
constrained quadratic program is solved at each Introducing vector notation, equations (2) and (3)
iteration through the p{ - & loop in Fig. 1. Hence, may be written in a more compact form as shown
the iterative loop guarantees that the control ac- below
tion calculated by the Volterra controller satisfies
the constraints on the manipulated and controlled y=Gu+c+f, (4)
variables. The model equations for the SISO case c = Hupast+ d + g. (5)
are given below
Equations (4) and (5) are similar to the equations
a1 0 *.. 0 in Ricker (1985) except for the presence of the f and
a2 a1 ‘. 0
g terms which arise due to the second-order model.
c is the invariant contribution of past inputs and
‘. al does not change with iterations around the p: - 4
= loop in Fig. 1. G and H are matrices that contain
. ‘. aI +a2
impulse response coefficients. upastis an historical
.. :
P-M+1
record of past values of the input, and u is the linear
up ap-1 ‘**
2 ai
I contribution of the present and future values of the
input. In addition, a setpoint trajectory, s, may be

I[’
c(k + II f(k+ I) implemented via the following equations:
c(k + 2) f(k + 2)
+ -I- (2)
s(k + 1) = hymeas
+ (I - A)~,,, (6)
[: c(k + P) /(k+ PI s(k+i)=hs(k+i-l)+(l-h)y,,,
21isP: (7)
I- a2 aj . +f ... a# 0
aj a4 ... cl&I 0 0 where h is a tuning parameter (in addition to M
and P) used to filter the setpoint.
To illustrate the computation off and g, consider
zz
ap_, ap : : : : the SISO case where M = 2, P = 3, and N =
I 4. Define B, a matrix containing the second-order
ape. i ii
coefficients, as
0 0 f 0 00 J

1
h,l h.2 h.3 h.4

0 b2.2 bz.3 b2.4


B= (8)
(3) 0 0 b3,3 b3.4
0 0 0 h.4

The bi,j terms represent coefficients in a triangular


P, M, and N are the output prediction horizon, in- Volterra model (equation (1)). At time k, u(k) and
put move horizon, and model truncation order, re- u(k + 1) are calculated. The entries in fare given by
spectively. The future-future and future-past cross
terms of the input correspond to f(k + 1), f(k + j-(k + 1) = [u(k) 0 0 O]B
1288 B. R. Maner et al.

x[u(k) u(k - 1) u(k - 2) u(k - 3)lT, 2.2. Nonlinear programming formulation


(9) The nonlinear programming formulation re-
quires solving the following optimization problem
f(k + 2) = [u(k + 1) u(k) 0 O]B
at each sampling interval:
x[u(k + 1) u(k) u(k - 1) u(k - 2)lT,
P
(10) m;ln c yh(k + i) - y(k + i))*
f(k + 3) = [u(k + 2) u(k + 1) u(k) O]B i=l

x[u(k + 2) u(k + 1) u(k) u(k - l)lT.


+ f h&Mk+j- l))*
(11) j=l

subject to
Since M = 2, u(k + 1) = u(k + 2) in equation (11).
The entries in g are given by N
y(k + i) = 2 alu(k + i - 2)
!?=I
g(k+l)=[O u(k-1) u(k-2) u(k-3)]B
N N
x[O u(k - 1) u(k - 2) u(k - 3)lT, + 1 b!,+(k + i - C)u(k + i - m),
1
e=1m=d
(12)
g(k+2)=[0 0 u(k-1) u(k-2)]B
u= [u(k) u(k+ 1) u(k+M- l)]q
x[O 0 u(k- 1) u(k- 2)lT, (13)
(18)
g(k + 3) = [O 0 0 u(k- l)]B
x[O 0 0 u(k- l)lT. Wow 5 u 5 Uhighp
(14) AUiowI AU I Atthigh.
For the unconstrained case, the control action is The approach in Section 2.1 involves a successive
computed at time k using the following algorithm: substitution approach to solve the optimization
Step 1. Set i= 1. problem in equation (18). An alternative method
Step 2. Calculate a and II by solving the uncon- to calculate the control action would employ a gen-
strained least squares control problem eral nonlinear programming solver to directly solve
the optimization problem. It should be noted that
a=((~-c-f)~G)~, equation (18) does not include output constraints.
(15)
Their inclusion would lead to a more complex
u = (GTG)-‘a. (16) nonlinear programming problem involving non-
linear constraints as well as a nonlinear objective
Step 3. Determine if the condition in equa- function.
tion (17) is satisfied where b is the desired
tolerance
3. CALCULATION OF MODEL PARAMETERS FOR A
1#j(k) _ &-l) MULTIVARIABLE CASE
(k) I I b (17)
The multivariable equation for output i having q
and the superscript denotes the iteration inputs for the second-order Volterra model used in
step. this work is given by
Step 4.. If yes, set u(k) = di)(k). Close the
Y N
switch in Fig. 1 and implement u(k).
If no, recalculate f using au)(k) for
Yitk) = yio + C1 af,ju,(k - j)
/=I j=l
present and future values of the input. Y N N
Set i = i + 1. Return to Step 2. (For
the constrained case, u is calculated
+ 1 C 1 bi,j,,,P/tk- Au/ tk - n),
/=I j=ln=j

by solving a quadratic program rather


(19)
than using the least-squares solution in
equation (16).) i.e. no cross terms in the inputs were permitted.
Contraction mapping arguments can be made to The yis term in equation (19) is a bias term whose
show convergence of the successive substitution al- value could be fitted using input-output data. How-
gorithm (Economou, 1985). Additional details of ever, no bias term arises when an analytical con-
the algorithm as well as examples of the constraint struction of the model parameters is used. To il-
handling and disturbance rejection capabilities ap- lustrate the procedure for the multivariable case, a
pear in Maner (1993). two-input-two-output system will be considered. A
Volterra MPC of polymerization reactors 1289

continuous-time bilinear model is obtained from a The resultin discrete-time model matrices for the
fundamental model using Carleman linearization bilinear mo %el are
(Rugh, 1981)

i(t) = AZ(~) + N,z(t)ur 0) + N2z(t)u2(t) + Bu(t)


A-I+;
1
;h3A4+h2A3+3hA2+6A
1, (32)

Ri s g [ $h3 (A31Qi+ A2RiA + ARiA’ + fiiA3)]


= f(z(t), u(t)), (20)
+i [h2 (A2fii + AfiiA + RiA2)]
y(t) = Cz(t). (21)
+ i [3h (ARi + &A) + 614/l, i= i,2, (33)
Equations (20) and (21) describe the resulting bi-
linear model where the variables, u, x, and y are 8”’ E ; +h3A3B + h2A2ti + 3hAh + se], (34)
in deviation form, and the state vector, x, contains 1
the original states (XI, x2, . . .) and the second-order Bt2’(:, i) I a [ $h3 (A2RiB(:, i) + ARiAB(:, i) + 1QiA2h(:, i))]
cross terms of the states (4, ~1x2, . . .). The inputs
and outputs are scaled to produce u;,~ and b;,j,,,n + 5 [h2 (A&‘#(:, i) + RiAB(:, i)) + 3hIQiB(:.i)],
parameters that yield well-conditioned matrices for i= 1.2. (35)
use in nonlinear MPC. Poor scaling can lead to
singularity and roundoff problems which would be The Bt2) matrix arises in the derivation of the dis-
problematic for a nonlinear programming solver. cretization method. BC:,i) denotes the ith column
Diagonal scaling matrices were introduced for the of the matrix, B. The discrete-time model is then
inputs and outputs given by

z(k + 1) = AZ(k) + &z(k)ur (k)


Nt= (22)
+&z(k)zQ(k) + B”‘U(k)
+B’2’(:, 1)&k) + B@)(.., 2)&k) ,
(23) (36)
y(k) = &z(k). (37)
and the state space matrices were redefined as fol-
lows: A derivation of discrete-time matrices using the
improved Euler method and relationships between
fii = NiN,,, i= 1,2, (24) discrete-time and continuous-time bilinear model
B = BN,, matrices are given in the Appendix of Maner (1993).
(25)
This method yields satisfactory results for bilinear
&=N;‘C. (26) systems that have a low degree of stiffness
In this work, the scaling factors were chosen to be The parameters for a second-order, triangular
approximately equal to the maximum range of each Volterra model are calculated using the following
input and output. equations (Rugh, 198 1):

1
A fourth-order explicit Runge-Kutta scheme is
used to obtain a discrete-time bilinear model. Since ui,i d.i = eA’-‘B(‘), i 2 1,
(38)
the full nonlinear equations are integrated over a [ aLi a22.i
time interval while holding the manipulated vari-
i 2 1, (39)
ables constant, u was treated as a constant vec-
tor in the derivation of the discretization scheme.
One fourth-order Runge-Kutta method with dis- = ~~j-*~lAi-j-lBcI)(. 1)
.e ,
cretization step size h uses the following equation
(Carnahan et al., 1969): jl l,ilj+ 1, (40)

z(k+ 1) =z(k) + ;(k, +2k2+2k3+&),


(27)
b!!,j,2.i
[ b:j,2.i 1 = ~dj-‘fi2~‘-j-lfj(‘)(a

jl l,irj+ 1.
21,

(41)
where In equation (38), aii corresponds to the jth im-
pulse response coefficient relating input I to output
kr = f(z(k)), (28) i* bf.j.l.n corresponds to the second-order Volterra
parameter with delays j and n relating input 1 to
k2 = f(z(k) + $hkl), (29)
output i. The bi,j,,,n parameters are not included,
since the contribution of these terms did not justify
k3 = f(z(k) + ;hk2), (30) the increased complexity that would be obtained if
kq = f(z(k) + hk3). (31) they were added to the model.
1290 B. R. Maner et al.

4. SIMULATION RESULTS
Table I. Parameters for Case Study I
4.1. Case Study I: SISO control of an isothermal
kr, = 1.3281 x 1O’O mJ/(kmol h)
CSTR 1.0930 x 10” &/(kmol h)
kxj =
As a first example, consider the isothermal free- kr = 1.0225 x lo-’ l/h
radical polymerization of methyl methacrylate us- k; = 2.4952 x IO6 m3/(kmol h)
ing azo-bis-isobutyronitrile as initiator and toluene h,, = 2.4522 x IO’ mg/(kmol h)
/* = 0.58
as solvent (Congalidis et al., 1989; Daoutidis et al.,
F = 1.00 d/h
1990; Doyle III et al., 1995). The number average v = 0.1 m3
molecular weight (NAMW) is controlled by ma- 4i. = 8.0 kmol/&
nipulating the inlet initiator flow rate. A schematic Mm = 100.12 kg/km01
of the process is shown in Fig. 2. The isothermal cm., = 6.0 kmol/mj
model was obtained by setting the reactor temper-
ature equal to its steady-state value of 335 K. Un- Table 2. Steady-state operating conditions for Case
der these assumptions, the six state model in Daou- Study I
tidis et al. (1990) reduces, to the following four state
model: XI =c, = 5.506114 kmol/m-’
x2 = q = 0.132906 kmol/m“”
dG F(G;v GJ), (42)
x3 =Do = 0.0019752 kmol/m3
- = -(kr + kr,)cmPo + _r4=D1 = 49.38182 kg/m3
dt u=fi = 0.016783 m3/h
dG JiCI, - J’CI =
” 25.000.5 ka/krrn’
- = -/kg1 + (43)
dt ’ v
dDo
- = (0.5/k& + kr,,)P, + kf,,C&$ - ?.(44) tual nonlinear model integrated with a fourth-order
dt Runge-Kutta method.
dDi FDi Open-loop responses of the nonlinear, linear,
-=A&(k,+k&P0-y, (45)
dt and second-order Volterra models to step changes
y=$. (46) of kO.008392 m”/h in F, from its nominal value of
0.016783 m3/h are shown in Fig. 3. In both sim-
where ulations, the output of the second-order Volterra
model more closely tracks the actual nonlinear

1
2f *kIcI o.5 plant output. The output of the Volterra model
PO
= [ kT, + kr, ’ is asymmetric for steps of the same magnitude of
opposite sign, while that of the linear model is
The model parameters and steady-state operating symmetric.
conditions are listed in Tables 1 and 2, respec- The result of a closed-loop simulation for a
tively. The variables 11, x, and y were placed in setpoint change from 25,000.5-38,000 kg/km01
deviation form and scaled by their nominal values is shown in Fig. 4. The reference trajectory was
(j&GQ~i==iJr and J = y), and a bilinear calculated by performing a closed-loop simulation
represiitation o?tde nonlinear model was ob- with the same tuning parameters using a linear
tained using Carleman linearization (Rugh, 198 1). plant and assuming no plant-model mismatch. The
The continuous-time bilinear system was then Volterra controller effectively cancels the process
discretized using the explicit fourth-order Runge- nonlinearity to produce a linear response. If all of
Kutta method given by equations (32)-(35) with a the nonlinearity was canceled by the controller, the
sampling time of 0.03 h (1.8 min). The largest neg- closed-loop response would be identical to the ref-
ative eigenvalue of the system was - 10.9 h-r. The erence trajectory. In this simulation, the response
corresponding small time constant for this system obtained by the Volterra controller tracks the ref-
arises from the relatively small reactor volume of erence trajectory more closely than the response
lOO+?.The model parameters were calculated using obtained by the linear controller. Linear MPC
equations (38)-(40). Linear MPC and the Volterra leads to an overshoot in the setpoint response,
controller described in Section 2.1 were used with and this may be undesirable if product specifica-
A4 = 1, P = 25, and h = 0.95 as tuning parame- tions require the molecular weight to be less than
ters. The tuning parameters were the same for both 38,000 kg/kmol. The cause of the overshoot can be
controllers because the underlying linear models attributed to the aggressive manipulated variable
are identical. The truncation order of the model profile for linear MPC shown in Fig. 4. While lin-
was N = 30. ear MPC could be detuned to yield an overdamped
This value was chosen to obtain a model with response for this setpoint change, detuning would
a memory approximately equal to three times the cause performance deterioration for setpoints be-
time constant of the system. The plant was the ac- low the nominal value of 25,000.5 kg/kmol. Hence,
Volterra MPC of polymerization reactors 1291

Monomer and Solvent

F cI D1

Cm T Dg

Fig. 2. Control configuration for Case Study I.

NAMW

1
\

.;I
._ ---_
2- -~--------------------__

1.8

t
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Time [hr]

Fig. 3. Open-loop simulations for step changes of kO.008392 m3/h in FI from its nominal value of
0.016783 m3ih. Solid: nonlinear; dashed: linear; dotted: Volterra.

the nonlinear behavior of this process requires a controller. Results for other setpoint changes are
compromise in the tuning of a linear model-based listed in Table 3.
1292 B. R. Maner et al.

2.5 1 I I I I 1 I
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Time [hr]

0.01 -
7; i’f.1, L,
[m3ihr] Li 1.1, - -.- _._,
o,oo5 ;;A ‘.., _l.T,-.~r__-._._._.__.-
- -.-.-.m._ ._._._._,_
._._._._._._ %=:=~-~---,-,-‘---I:
LL_ _--
.I
0’
I I I I I a I I I I
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Time [hr]

Fig. 4. Closed-loop simulation for a filtered step setpoint change from 25000.5 to 38,000 kg/kmol. Solid:
reference; dashed: linear MPC; dotted: Volterra MPC; dash dot: nonlinear QDMC.

However, the initial chattering observed in the ini-


Table 3. Performance comparison of the two con-
trollers for Case Study I tiator flow rate using nonlinear QDMC indicates
that this controller is already tuned quite aggres-
Setpoint Percent max. # of
sively. The poor performance is due to the difficulty
kg/km01 improvement in ISE iterations
84.79 27
that nonlinear QDMC encounters with large input
38,000
32,000 93.23 8 variable changes. Since the linear model employed
28,000 91.29 4 in linear MPC and the linear model used by non-
22,000 74.41 4 linear QDMC at time zero are identical, the first
18.000 unstable NA input move is the same for both algorithms. This
12.000 unstable NA
move corresponds to a change of -87.53% from the
nominal input value. The states in a physical non-
For a comparison with another nonlinear MPC linear model would not be expected to change dras-
strategy, the dash dot lines in Fig. 4 correspond tically in one sampling period. Hence, the values of
to input and output profiles obtained using non- the states after one sampling period could be ex-
linear quadratic dynamic matrix control (QDMC) pected to be close to their initial values. Prior to cal-
(Garcia, 1984) with M = 1, P = 5,and a first-order culating a second control move, nonlinear QDMC
filter in the feedback path with filter time constant linearizes the nonlinear model at the values of the
0.94. The tuning parameters for nonlinear QDMC input and states after one sampling interval. How-
were different from those used for linear MPC and ever, the linear model can be poor if these values do
Volterra MPC, because the underlying linear mod- not correspond to steady-state values. This would
els used by nonlinear QDMC were different from occur, for example, if a large input move was com-
those of the other two controllers. The tuning pa- puted and the states in the nonlinear model have
rameters used in nonlinear QDMC resulted in an not had time to respond to this input. This draw-
improvement in closed-loop performance as com- back of nonlinear QDMC was noted in the paper
pared to implementing the tuning parameters used presented by Garcia (1984) in that the algorithm ap-
in linear and Volterra MPC. In this implementa- proaches the performance of a rigorous nonlinear
tion, the actual nonlinear model is linearized at controller that solves a general nonlinear program-
each sampling interval, i.e. there is no plant-model ming problem “... for small input variable changes
mismatch. The output profile is much more sluggish and disturbances that do not move the operating
than the output trajectories obtained using linear point far.” For Case Study I, nonlinear QDMC re-
and Volterra MPC. It might be argued that non- peatedly obtains poor linear models at the begin-
linear QDMC should be tuned more aggressively.
Volterra MPC of polymerization reactors 1293

ning of the simulation as evidenced by the chatter- problem. To accomplish these objectives, the initia-
ing in the initiator flow rate profile. tor and cooling water flow rates were selected as
It is interesting to note the effect of reducing the manipulated variables. A schematic of this process
number of second-order coefficients on the closed- is shown in Fig. 9. The monomer flow rate was held
loop performance. Plots of the first and second- constant. The second modification to the original
order parameters are shown in Figs 5 and 6, re- problem is that the reactor is controlled around the
spectively. Since the bi,j coefficients are one order of low conversion stable steady-state point. Derivation
magnitude smaller than the ai coefficients in equa- of a bilinear model at a stable steady state enabled
tion (l), and ii is on the order of 10d3, the later bi,i calculation of the model parameters using equa-
terms in Fig. 6 make a very small contribution to tions (38)-(41). The nonlinear model was obtained
the model. Since N = 30, there are 465 bi,j terms. by augmenting the original four state model with
If the truncation order for the second-order param- two additional equations that are used to determine
eters is reduced from N = 30-15, there would be the number average molecular weight
120 bi,j terms. The closed-loop performance using
this reduced model is shown in Fig. 7. As expected, d[Zl
-= (Qi[hl - Qt[Il) _ kdiIl
(47)
the performance is very nearly the same as that
shown in Fig. 4 for the full second-order Volterra d&l _ (QnJM: - QtWl) _ k IMlkPl
P
model. This model structure pruning approach cor- dt V
responds to reducing the nonlinear memory of the (49
system as noted in Pearson et al. (n.d.). dT
-= Qt(fi- 5’3+ (-AH,,, [MIIPl
For setpoints below 18,000 kg/kmol, the closed- dt V PC, p
loop performance of the Volterra controller is
unstable while the linear model-based controller -hA(T - T,),
PCpV
successfully brings the nonlinear plant to the new
dT, QcWcf - T,) + hA
setpoint. An explanation for this behavior can be pCV(T - T,), (50)
found from a plot of the steady-state gain loci
dt= v, cpcc
(Fig. 8). The gain of the bilinear model changes dDo
- = O.Sk,[lq - y, (51)
sign in the region of Y = 18,000 kg/kmol. This dt
error is the result of deriving the bilinear model dD1
- = M,k,,[M][P] - 9, (52)
using local expansion results such as Carleman dt
linearization. This technique is accurate for captur-
(53)
ing local nonlinearities around an operating point,
but may be erroneous in describing global non- Y* = T. (54)
linear behavior (Doyle III et al, 1995). Since the
discrete-time Volterra model was obtained from the where
continuous-time model, it also erroneously predicts
a sign change in gain near Y = 18,000 kglkmol.
[p] = [ By]“.5,
The Volterra series is a time-invariant operator ki = Aiexp(-Ei/ T), i = d,p,t,
and would need parameter adaptation if it is to
be used over a larger region of operation. Several Qt = Qi + Qs+ Qm.
researchers (HernBndez and Arkun, 1993; Ricker Two assumptions were made in deriving equa-
and Lee, 1995) have incorporated on-line parame- tions (51) and (52). First, it was assumed that the
ter adaptation into nonlinear MPC schemes, and rate of disappearance of monomer was primar-
this is a promising direction for future research. ily due to propagation. Hence, disappearance of
monomer due to chain transfer to monomer was
not included in the model. This assumption was
4.2. Case Study II: MIMO control of a CSTR also made in Hidalgo and Brosilow (1990) and
The second case study considers the control of Jaisinghani and Ray (1977). The second assump-
the free-radical solution polymerization of styrene tion was that the overall chain termination rate
in a jacketed CSTR. Hidalgo and Brosilow (1990) constant, k,, was composed of both combination,
proposed a nonlinear MPC scheme to control the kT,> and disproportion?tion, kT, , contributions
reactor temperature at the unstable steady state by (Schmidt and Ray, 198 1) where
manipulating the cooling water and monomer flow
kl = kT, + kTd. (55)
rates. For this example, two changes were made to
the original problem. First, it is desired to control For styrene, chain termination has been experimen-
the number average molecular weight in addition to tally determined to occur solely by combination in
the reactor temperature to obtain a MIMO control bulk (Bevington et al., 1954) and solution (Timm
1294 B. R. Maner et al.

-0.025 -

Fig. 5. First-order coefficients for Case Study I.

x103

Fig. 6. Second-order coefficients for Case Study I.

and Rachow, 1974) polymerization. Hence, kt = tained at 0.6 so that the gel effect may be neglected
kT, in equation (55). The kinetic and thermody- (Choi, 1986). Hence, the solvent flow rate was ma-
namic parameters (which correspond to Process 1 nipulated according to the following equation:
in the original paper (Hidalgo and Brosilow, 1990)),
model parameters, and steady-state operating con- Qs = 1.5Qm - Qi. (56)
ditions are listed in Tables 4-6, respectively. The
model assumptions affecting the parameter selec- A second model assumption is that the temperature
tion are the same as those in the original paper must be less than 423 K. This is justified by the fact
(Hidalgo and Brosilow, 1990). The first assump- that the rate of thermal initiation becomes signifi-
tion is that the solvent volume fraction is main- cant at temperatures greater than 373 K, and prob-
Volterra MPC of polymerization reactors 1295

-_
NAMW

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.6 2


Time [hr]

0.015 -

Fl

0.01 -
r’ i!
[m?hr] 1’ ,’ 1.!iL.,._,
.! i ,! .: 1 ,, ,, -‘-._,_,
o.oo5 ;;a ..,,~~,~~-T. ,_, -‘-.-(-.-..w.-._____,_,_,
_.____.-- ._.__._.__._._=. ‘=.A.x.--~-.--^-z
.LL _ r- -
.I
0 I I 1 I I I
0 0.2 0.4 0.6 0.6 1 1.2 1.4 1.6 1.8 2
Time [hr]

Fig. 7. Closed-loop simulation for a liked step setpoint change from 25000.5 to 38.000 kgkmol using
the reduced model. Solid: reference; dashed: linear MPC; dotted: Volterra MPC; dash dot: nonlinear
QDMC.

lo2
NAMW

Fig. 8. Gain loci for various models. Solid: nonlinear; dashed: linear; dotted: Volterra; dash dot: bilinear.

ably dominates catalytic initiation at temperatures weights while keeping the temperature at its set-
above 423 K (Biesenberger and Sebastian, 1983). point. The vectors u, x, and y were placed in de-
The goal of the control system is to drive the viation form, and a bilinear representation of the
polymerization system to a new state to produce nonlinear model was again obtained using Carle-
polymers with different number average molecular man linearization (Rugh, 1981). The model matri-
1296 B. R. Maner et al.

Qs Tf
Solvent

Q, Wfl Tf

I Monomer

Qc Tc
Cooling fluid

Cooling fluid
I
. . . .._.... . .._....___ .._ .._
Q [Ml [II T
Effluent

Fig. 9. Control configuration for Case Study II.

Table 4. Kinetic and thermodynamic


for Case Study II
parameters
N =
Y
[
0 1
2500 0
0.5 . (58)

The scaled, continuous-time bilinear system was


Ad 5.95 x 10” I/S
discretized using the explicit fourth-order Runge-
Ed 14,891 K
At 1.25 x lo9 U(mol s) Kutta method given by equations (32)-(35) with a
Et 843 K sampling time of 1 h. Although the sampling inter-
AP 1.06 x 10’ l/(mol s) val is large, the sampling time is still less than one-
EP 3551 K
fifth the smallest time constant in the linear trans-
-AH, 16,700 cal/mol
hA 70 cal/(K s) fer function matrix. The open-loop and closed-loop
PC, 360 cal/(K 4?) responses are also of the same order as those ob-
966.3 cal/(K d) served in the control of a lOOOf?polymerization re-
-_.Es--
actor in Congalidis et al. (1989). The largest nega-
Table 5. Parameters for Case Study II tive eigenvalue of the Hidalgo and Brosilow system
was -0.74 h-r at the low conversion steady-state.
Qs = 0.1275 .%
The large time constant for this system arises from
Q,,, = 0.105 8/S

v = 3000 d the large reactor volume of 30004 and operation of


vc = 3312.4 d the reactor at the low temperature, low conversion
[IfI = 0.5888 moW steady state. The model parameters were calculated
[Mfl = 8.6981 mol/# using equations (38)-(41). The Volterra controller
rf = 330 K
described in Section 2.2 was used in this case study,
T,r = 295 K
Mm = 104.14 g/m01 i.e. a fourth-order nonlinear program was solved at
each sampling interval. The tuning parameters for
ces were scaled with the following scaling matrices: linear and Volterra MPC were A4 = 1, P = 20,
[yl, y4=[2,1], and [AI, h+[lOOO,l]. The truncation
order of the model was N = 35. The truncation
NU= [ ‘:” 4,9.6]’ (57) order was chosen to obtain a model with a mem-
Volterra MPC of polymerization reactors 1297

ory approximately equal to three times the largest The performance obtained using nonlinear
time constant of this process. The plant was the QDMC is comparable to that of using Volterra
actual nonlinear model integrated with the Gear MPC. The response of the number average molec-
predictor-corrector method. ular weight is approximately first order, and the
temperature is maintained close to its setpoint.
Table 6. Steady-state operating conditions for Case In addition, the input profiles for both flow rates
Study II are very smooth. The reason for the good perfor-
mance of nonlinear QDMC in this case study is
x1 = [I] = 6.6832 x lo-* mol/#
that the first move in initiator flow rate is - 18.19%)
x2=[Ml = 3.3245 molM
x3=T = 323.56 K from its nominal value, and the first move in cool-
xq = Tc = 305.17 K ing water flow rate is - 18.55% from its nominal
x5 = Do = 2.7547 x 1O-4 molM value. Hence, the input variable changes are much
.X6=& = 16.110 de smaller, on a percentage basis, than those encoun-
ul=Qi = 0.03 BlS
tered in Case Study I. The linear models obtained
uz = Qc = 0.131 7%
= 58,481 g/m01 at each sampling interval are more accurate when
Yl
Y2
= 323.56 K small input variable changes are used, and the cor-
responding control action is more appropriate. An
Open-loop responses of the nonlinear, linear, attractive feature of nonlinear QDMC is that a sin-
and second-order Volterra models to step changes gle quadratic program needs to be solved on-line,
of *27&h in Qj from its nominal value of 108$/h while the second-order Volterra controller requires
are shown in Fig. 10. In both simulations, the the solution of a fourth-order nonlinear program.
output of the second-order Volterra model more However, the nonlinear QDMC algorithm requires
closely tracks the actual nonlinear plant output the availability of a fundamental model. In ad-
for the number average molecular weight (and the dition, the nonlinear QDMC results presented in
curves are indistinguishable for one of the step both case studies assume there is no plant-model
changes). Although these step changes correspond mismatch, which will not be the case in practice.
to ?25%, changes in the initiator flow rate, the The simulation in Fig. 11 used a triangular
temperature does not significantly deviate from the second-order Volterra model that did not include
nominal value. Both the linear and second-order cross terms (ul(k - l)uz(k - l), . . ., etc.) in the
Volterra models accurately predict the moderate inputs. For a truncation order of N = 35, this re-
temperature changes of the plant. sulted in 2520 second-order coefficients. The first
The result of a closed-loop simulation for a set- and second-order parameters for this case study
point change from 58,48 1 to 80,000 g/mol is shown are shown in Figs 12 and 13, respectively. The first-
in Fig. 11. As in the preceding example, the Volterra order parameters are approximately one order of
model predictive controller outperforms the linear magnitude larger than the second-order terms, and
model predictive controller, as the number aver- the scaled manipulated variables were on the order
age molecular weight more closely tracks the refer- of 10-l. Hence, the later b;,j,,,, terms in Fig. 13
ence trajectory. Conversely, linear MPC results in make a very small contribution to the model. Al-
significant overshoot. The overshoot is caused by though the reactor temperature is affected by the
the more aggressive manipulated variable profiles heat of polymerization and the Arrhenius temper-
in Fig. 11. Linear MPC may be detuned to yield ature dependence of the kinetic parameters, the
an overdamped response for the number average large reactor acts as a heat sink and prevents a large
molecular weight. However, improved performance change in temperature provided there is always a
for positive setpoint changes would be achieved at nominal amount of cooling water flowing through
the expense of performance deterioration for nega- the cooling jacket. Hence, linear models could be
tive setpoint changes. used to describe the relationship between the two
The dash dot lines in Fig. 11 correspond to in- inputs and the reactor temperature. In addition,
put and output profiles obtained using nonlinear the initiator flow rate has a much greater effect
QDMC with A4 = 1, P = 12, [ri, n]=[1,2000], on the number average molecular weight than the
and [hi, hz]=[O,O].Different tuning parameters were cooling water flow rate, and a linear model could
used for nonlinear QDMC as compared to those also be used to relate the cooling water flow rate
used for linear and Volterra MPC, because the un- and the number average molecular weight. In ad-
derlying linear models were different from those dition, if the truncation order of the second-order
used in the other two controllers. The tuning pa- parameters in the remaining Volterra model is re-
rameters used in nonlinear QDMC resulted in an duced from N = 35-20, using the same reasoning
improvement in closed-loop performance as com- in the first case study, the number of second-order
pared to incorporating the tuning parameters used coefficients is reduced to 210. Repeating the same
in linear and Volterra MPC.
1298 B. R. Maner et al.

4.5c , I 4
0 10 20 30 40 50 60
Time [hr]

3221 , , , , , j

0 10 20 30 40 50 60
Time [hr]

Fig. 10. Open-loop simulations for step changes of &27&I in Qi from its nominal value of 108Hh. Solid:
nonlinear; dashed: linear; dotted: Volterra.

:jI-.~~.~T.:
....
318
318.
I 314’ I
20 40 60 0 20 40 60
Time [hr] Time [hr]

500

100.
400.
Qi Q,
,i .‘.., f.J... ,,.
~\~~~~~~.-..-.---.-
[Whr] 80:i ,, Wrl 300 .‘!, /.”
60. \.., .y\z;;.-,.u-____--.
:. 7 .--
200: 1’

j\/ . 1 loo-
/

0 20 40 60 0 20 60
Time [hr] Time [hr]

Fig. II. Closed-loop simulation for a step setpoint change from 58.481 to 80,000 g/mol. Solid: reference;
dashed: linear MPC; dotted: Volterra MPC; dash dot: nonlinear QDMC.

setpoint change simulation yielded the closed-loop second-order Volterra models require many pa-
performance illustrated in Fig. 14. The closed-loop rameters to describe nonlinearities. In both case
performance does not signticantly degrade with studies, it was shown that the number of second-
the elimination of over 900/n of the second-order order coefhcients could be significantly reduced
parameters in the second-order Volterra model. with no significant difference in closed-loop perfor-
One of the drawbacks of this approach is that mance. While it seems reasonable that the memory
Volterra MPC of polymerization reactors 1299

-2
0 10 20 30 40 0 10 20 30 40
i i

O-

-0.2

-0.4.

-0.6.

-0.6.f m

-1
0 10 20 30 40
i

Fig. 12. First-order coefficients for Case Study II.

i 00 j I 00 j

0.05

b2 1.j.l.l
0

i 00 j

Fig. 13. Second-order coefficients for Case Study II.

of the linear and second-order contributions to second-order nonlinear model. The performance of
the system dynamics should be approximately the a linear model-based controller usually degrades as
same, the bi,j,,,, parameters are one order of mag- the system is moved away from the point at which
nitude smaller than the ai,j coefficients, and act the model was obtained. Although performance
to correct the linear model for the effect of non- degradation occurs, the linear model-based con-
linearities. Hence, the later b%j,r,,terms are very troller is still stable over a fairly broad regime. The
small and may be omitted without significant per- performance of a second-order nonlinear model-
formance degradation in modeling and control. based controller also typically worsens as the plant
Another disadvantage is associated with using a is moved away from the point at which the model
1300 B. R. Maner et al.

x104
8.5 _
I \
326 71
324
322
320
318
316
I--=-
5.5 I
0 20 40 60 20 40 60
Time [hr] Time [hr]

100.
Qi QC
80 .I
VW Ifir1
so. ;!_:1-‘” _._..-.-_ _ _.__.

0 20 40 60
Time [hr]

Fig. 14. Closed-loop simulation for a step setpoint change from 58,481 to 80,000 g/mol using the reduced
model. Solid: reference; dashed: linear MPC; dotted: Volterra MPC; dash dot: nonlinear QDMC.

was obtained. However, due to the inherent cur- and Nikolaou, 1995) have been presented by other
vature of second-order models, they will always researchers. In addition, the ability to obtain the
predict that the plant gain will change sign at some second-order Volterra model parameters from Car-
point. Hence, care should be taken to ensure that leman linearization and input-output data is an-
a second-order model-based controller is not used other advantage.
outside the region for which the model is valid. The
ability of a second-order Volterra model derived
from Carleman linearization to capture ‘local’
5. CONCLUSIONS
and not global behavior was also demonstrated in
Doyle III et al. (1995). The ability of second-order Volterra models
Despite the drawbacks associated with the pro- to capture asymmetric changes in output given
posed second-order Volterra controller, there are symmetric changes in input was illustrated with
also several advantages of this control scheme. One open-loop simulations. A nonlinear MPC scheme
advantage is that improved performance over a lin- utilizing a successive substitution algorithm was
ear model-based controller can be achieved as il- presented. This controller yielded improved perfor-
lustrated in both case studies. A second advantage mance over linear MPC for the control of number
is that these models yield a fourth-order nonlin- average molecular weight in an isothermal CSTR.
ear program for the standard two-norm MPC ob- Improved multivariable control of the number av-
jective function if the controller is implemented as erage molecular weight and temperature in a large
shown in Section 2.2. While this optimization prob- polymerization reactor was also demonstrated.
lem is more difficult to solve than the quadratic pro- The models were obtained by Carleman lineariza-
gram encountered using traditional linear MPC, it tion of the fundamental model, discretization of
is less computationally burdensome than the non- the resulting continuous-time bilinear model, and
linear program that arises in modified nonlinear in- calculation of the parameters for a triangular
ternal model control or nonlinear MPC using poly- second-order Volterra model. Relationships be-
nomial ARMA models. A third advantage results tween discrete-time bilinear model matrices and
from the fact that Volterra models are based only their continuous-time counterparts using an ex-
on past inputs. Hence, stability results are easier to plicit fourth-order Runge-Kutta method were also
obtain than for models based on past inputs and presented. It was also demonstrated that a large re-
outputs. Stability results for the continuous-time duction in the number of second-order parameters
inverse Volterra series (Zheng and Zafiriou, 1993) could be made without a significant degradation
and the one-norm SISO MPC formulation (Genceli in closed-loop performance.
Volterra MPC of polymerization reactors

Acknowledgements- FJD and BRM would like to acknowl- Hidalgo, P M. and C. B. Brosilow (1990). Nonlinear model
edge funding from the National Science Foundation NY1 pro- predictive control of styrene polymerization at unstable op-
gram (Grant CTS-9257059). Additional support for BRM was erating points. Computers Chem. Engng, 14(4/5), 481-494.
provided by a Purdue University Andrews Fellowship. Jaisinghani, R. and W. H. Ray (1977). On the dynamic be-
haviour of a class of homogeneous continuous stirred tank
polymerization reactors. Chem. Engng Sci., 32, 811-825.
REFERENCES Lee, J. H. and N. L. Ricker (1993). Extended Kalman fil-
ter based nonlinear model predictive control. In Proc. Am.
Bevington, J. C., H. W. Melville and R. P. Taylor (1954). The Control Con$, San Francisco, CA. pp. 1895-1899.
termination reaction in radical polymerizations. II Polymer- Li, W. C. and L. T. Biegler (1988). Process control strategies
izations of styrene at 60° and of methyl methacrylate at 0 for constrained nonlinear systems. Znd. Engng Chem. Res.,
and 60°, and the copolymerization of these monomers at 27, 1421-1433.
60°. J. Poiym. Sci., XIV, 463476. Li, W. C. and L. T. Biegler (1989). Multistep, Newton-type
Biesenberger, J. A. and D. H. Sebastian (1983). Principles of control strategies for constrained, nonlinear processes. Chem.
Polymerization Engineering. John Wiley & Sons, New York. Engng Res. Des., 67, 562-577.
Carnahan, B., H. A. Luther and J. 0. Wilkes (1969). Appfied Maner, B. R. (1993). Nonlinear model predictive control with
Numerical Methods. John Wiley & Sons, New York. second order Volterra models. Master’s thesis, Chemical
Choi, K. Y. (1986). Analysis of steady state of free radical Engineering Dept, Purdue University, West Lafayette, IN.
solution polymerization in a continuous stirred tank reactor. Pearson, R. K., B. A. Ogunnaike and F. J. Doyle III (1996).
Polymer Engng. Sci., 26(14), 975-981. Identification of discrete convolution models for nonlinear
Chow, T., E. Eskow and R. Schnabel (1994). Algorithm 739: processes. IEEE Trans. Acoustics, Speech, and Signal Pro-
A software package for unconstrained optimization using cessing, in press.
tensor methods. ACM Trans. Math. Software, 20(4), 518- Pearson, R. K., B. A. Ogunnaike and F. J. Doyle III (n.d.).
530. Second-order Volterra models for chemical processes, Part
Congalidis, J. I?, J. R. Richards and W. H. Ray (1989). Feedfor- I. In preparation.
ward and feedback control of a solution copolymerization Ricker, N. L. (1985). Use of quadratic programming for con-
reactor. AIChE J, 35(6), 891-907. strained internal model control. Ind. Engng Chem. Process
Daoutidis, P, M. Soroush and C. Kravaris (1990). Feedfor- Des. Dev., 24, 925-936.
ward/feedback control of multivariable nonlinear processes. Ricker, N. L. and J. H. Lee (1995). Nonlinear model predictive
AIChE J., 36(10), 1471-1484. control of the Tennessee Eastman challenge process. Com-
Doyle III, F. J., B. A. Ogunnaike and R. K. Pearson (1995). puters Chem. Engng, 19(9), 961-98 1.
Nonlinear model-based control using second-order Volterra Rugh, W. J. (1981). Nonlinear System Theory - The
models. Automatica, 31(5), 697-714. VoiterraiWiener Approach. The Johns Hopkins University
Economou, C. G. (1985). An operator theory approach to non- Press, Baltimore, MD.
linear controller design. PhD thesis, Chemical Engineering, Schmidt, A. D. and W. H. Ray (1981). The dynamic behavior of
California Institute of Technology, Pasadena, CA. continuous polymerization reactors-I Isothermal solution
Economou, C. G. and M. Morari (1985). Newton control laws polymerization in a CSTR. Chem. Engng Sci., 36, 1401-
for nonlinear controller design. In Proc. IEEE Conf: on 1410.
Decision and Control, Fort Lauderdale, FL, pp. 1361-1366. Timm, D. C. and J. W. Rachow (1974). Description of polymer-
Garcia, C. E. (1984). Quadratic dynamic matrix control of non- ization dynamics by using population density. In H. M. Hul-
linear processes. An application to a batch reaction process. burt (Ed.), Chemical Reaction Engineering--II9 Advances in
AIChE Annual Meeting, San Francisco, CA. Chemistry Series 133. American Chemical Society, pp. 122-
Gattu, G. and E. Zafiriou (1992). Nonlinear quadratic dynamic 136.
matrix control with state estimation. Ind. Engng Chem. Res., Zames, G. (I 966). On the input-output stability of time-varying
31, 1096-l 104. nonlinear feedback systems, part I: Conditions derived using
Genceli, H. and M. Nikolaou (1995). Design of robust con- concepts of loop gain, conicity, and positivity. IEEE Trans.
strained model-predictive controllers with Volterra series. Automat. Control, AC-11(2), 228-238.
AIChE J., 41(9), 2098-2107. Zheng, Q. and E. Zafiriou (1993). Stability analysis of inverse
Hernandez, E. (1992). Control of nonlinear systems using Volterra series. Presented at the AIChE Annual Meeting, St.
input-output information. PhD thesis, Chemical Engineer- Louis, MO.
ing, Georgia Institute of Technology, Atlanta, GA.
Hernandez, E. and Y. Arkun (1993). Control of nonlin-
ear systems using polynomial ARMA models. AZChE J.,
39(3), 446-460.

Vous aimerez peut-être aussi