Sie sind auf Seite 1von 14

Chemical Engineering Science 59 (2004) 4325 4338

www.elsevier.com/locate/ces

A method for robustness analysis of controlled nonlinear systems


Juergen Hahn , Martin Mnnigmann, Wolfgang Marquardt
Lehrstuhl fr Prozesstechnik, RWTH Aachen, Germany
Received 11 March 2004; received in revised form 6 June 2004; accepted 16 June 2004

Abstract
This paper presents an approach to analyzing robustness properties of nonlinear systems under feedback control. The core idea is
to apply numerical bifurcation analysis to the closed-loop process, using the controller/observer tuning parameters, the set points, and
parameters describing model uncertainty (parametric as well as unmodeled dynamics) as bifurcation parameters. By analyzing the Hopf
bifurcation and saddle-node bifurcation loci with respect to these parameters, bounds on the controller tuning are identied which can
serve as a measure for the robustness of the controlled system. These bounds depend upon the type as well as the degree of mismatch
that exists between the plant and the model used for controller design.
The method is illustrated by analyzing three control systems which are applied to a continuously operated stirred tank reactor: a
state feedback linearizing controller and two output feedback linearizing controllers. While model uncertainty has only a minor effect
on the tuning of the state feedback linearizing controller, this does not represent a very realistic scenario. However, when an observer
is implemented in addition to the controller and an output feedback linearizing scheme is investigated, it is found that the plant-model
mismatch has a much more profound impact on the tuning of the observer than it has on the controller tuning. In addition, two observer
designs with different level of complexity are investigated and it is found that a scheme which makes use of additional knowledge about
the system will not necessarily result in better stability properties as the level of uncertainty in the model increases. These investigations
are carried out using the robustness analysis scheme introduced in this paper.
2004 Elsevier Ltd. All rights reserved.
Keywords: Nonlinear system; Bifurcation analysis; Feedback linearizing control; Robust controller tuning

1. Introduction
Over the past few years nonlinear process control has become increasingly popular in the chemical process industries. This development is due to the trend towards specialty
products, tighter prot margins, more stringent environmental requirements, as well as advances in nonlinear systems
theory and in the numerical implementation of nonlinear
controllers (Bequette, 1991).
Many different techniques for designing controllers for
nonlinear processes have been introduced in the literature.
One of the simplest methods is to linearize the process model
and to design a linear controller of PID type for the system.
Corresponding author. Present address: Department of Chemical
Engineering,Texas A&M University, College Station, TX 77843-3122,
USA. Tel.: +1-979-845-3568; fax: +1-979-845-6446.
E-mail address: hahn@tamu.edu (J. Hahn).

0009-2509/$ - see front matter 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2004.06.026

More complex controller design techniques take the nonlinear behavior of the process into account. This category
of methods includes feedback linearization and nonlinear
model predictive control. If an accurate model of the process is known, then controller designs that make use of additional knowledge about the process can offer performance
advantages. However, the larger the model uncertainty the
more this performance advantage will degrade, even up to
the point that the closed-loop system can become unstable.
The described scenario presents the trade-off between performance and robustness (Henson and Seborg, 1991).
This paper introduces a systematic approach to analyzing the effect of model mismatch on controller and observer
design and tuning for nonlinear systems over a region of
operation. Bifurcation analysis is applied to controlled nonlinear systems by using the controller and observer parameters, the set points of the controller, as well as parameters

4326

J. Hahn et al. / Chemical Engineering Science 59 (2004) 4325 4338

describing the mismatch between the plant and the model as


bifurcation parameters. This way it is possible to determine
controller/observer tuning parameters which result in stable
operation of the closed-loop system for a specied amount
of model uncertainty. It also has to be avoided that a computed stable operating point lies in the vicinity of an unstable operating region into which a trajectory of the process
might lead. This can be investigated by using the set points
of the control system as parameters for the bifurcation analysis. In particular, it is often possible to compute bounds for
the controller tuning parameters that guarantee the existence
of only stable operating points for a specied type/amount
of model mismatch over the entire operating region.
It should be noted that this bifurcation analysis-based approach will put specic emphasis on the effect that model
uncertainty has on controller tuning and less emphasis on determining an appropriate controller structure. However, the
size of the region of stability for the controller/observer tuning parameters can be analyzed for many different control
structures and can serve as a basis for comparing robustness
properties of different controllers for a specic system.
The procedure is illustrated via a case study comparing
three different controllers: application of a state feedback
linearizing controller and two output feedback linearizing
controllers to a continuously stirred tank reactor. For these
case studies, uncertainty in the model parameters and plantmodel mismatch due to unmodeled dynamics are investigated. Parametric uncertainty results in an upper bound on
the controller/observer tuning parameters for the feedback
linearizing controller. In contrast, unmodeled dynamics result in a lower bound for the tuning parameter for the observers and controllers. From bifurcation analysis it can be
inferred that the feedback linearizing controllers can guarantee stable operating points over the entire operating region
for values of the tuning parameters between these bounds.
The organization of the paper is as follows. Section 2 reviews background material on bifurcation analysis and controller/observer design. Specic emphasis is put on presenting the material in a manner consistent with the scope of this
work. This is especially important since bifurcation analysis and control theory are two areas of research that have
been developed independently from one another. Section 3
presents the main contribution of this work by describing
the ideas and methods behind applying bifurcation analysis
to controlled systems. A subsection describing how the proposed framework can be used for determining robust controller tuning parameters is included. The method is illustrated via case studies in Sections 4, and Section 5 contains
the conclusions.
2. Preliminaries
2.1. Bifurcation analysis
Numerical bifurcation analysis by continuation has been
used by the chemical engineering community for three

decades. A seminal example is the analysis of various models of the exothermic continuous stirred tank reactor (see
Uppal et al., 1974, for example). Textbooks on bifurcation
theory and bifurcation analysis by parameter continuation
exist that address the subject with various degrees of mathematical rigor (Seydel, 1994; Perko, 1996; Guckenheimer
and Holmes, 1993; Kuznetsov, 1998b). In addition, software
for bifurcation analysis by parameter continuation is available for dynamic systems of different size, most notably
auto (Doedel et al., 2000), content (Kuznetsov, 1998a), and
DIVA (Mangold et al., 2000).
Bifurcation analysis by continuation involves linearizations of generally nonlinear process models. This often
causes confusion as readers assume that linearizations inevitably imply that the analysis is local only. This section
points out how numerical bifurcation analysis by continuation can in fact be used for more than local analysis of
nonlinear systemsdespite using linearizations.
Assume that a nonlinear model of the form
x = f (x, u, p)

(1)

is given, where x, u and p are nx state variables, nu input


variables, and np parameters, respectively. Functions f map
from a subset of R nx R nu R np into R nx . These functions
are assumed to be smooth with respect to x, u and p. The
notation distinguishes between input variables u and process
parameters p. The reason for this distinction will become
evident in Section 3. However, for a continuation analysis it
is not natural to distinguish between u and p. In the present
section, input variables u and parameters p are therefore
abbreviated by
u = (uT , pT )T R nu ,

nu = nu + np .

(2)

Eq. (1) can now be rewritten as


x = f (x, u).

(3)

A rst step towards analyzing the stability properties of


a steady state 0 = f (xss , u ss ) of system (3) could be to
linearize it. This yields the approximation
x = A(xss , u ss ) (x xss ) + B(xss , u ss ) (u u ss ),

(4)

where A(xss , u ss ), B(xss , u ss ) denote the nx nx dimensional and nx nu -dimensional Jacobians of f w.r.t. x
and u,
respectively, evaluated at the steady state (xss , u ss ).
Clearly, linearization (4) will only be valid for sufciently
small ||x xss || and ||u u ss ||. Note in particular that Eq.
(4) will become invalid as the input u leaves the vicinity
of u ss .
As opposed to linearizing at a steady state, numerical bifurcation analysis by continuation uses linearizations along
curves of steady states. Assume that  : U R nx is a curve
of equilibria which is parameterized by one of the u i while
all remaining parameters are kept at xed values u j = u j,ss
j = i. Without loss of generality the curve  is parameterized by the rst parameter u 1 , and the remaining u 2 , ..., u nu

J. Hahn et al. / Chemical Engineering Science 59 (2004) 4325 4338

are xed in the sequel. The assumption that  is a curve of


equilibria implies

25

0 = f ((u 1 ), u 1 , u 2,ss , ..., u nu ,ss )

20

(5)

0 = f ((u 1 ), u 1 ).

(6)

In contrast to the linearization at a steady state given by


Eq. (4), the nonlinear model (3) can be linearized along the
curve ,



x = f ((u 1 ), u 1 ) = A (u 1 ), u 1 x (u 1 ) + h.o.t., (7)
where the xed arguments u 2,ss , ..., u nu ,ss have again been
omitted for simplicity. It is stressed that in numerical continuation, it is not necessary to know a curve  in advance.
It fact  need never be calculated explicitly, but it is built
up numerically as the analysis proceeds. This is illustrated
with an example given below.
While the difference between linearizations (4) and (7) is
technical at rst sight, it is crucial to use Eq. (7) in a bifurcation analysis by continuation, if more than local information
on system stability is to be obtained. In order to infer the
stability of the nonlinear system (3), the following basic, yet
powerful, result from nonlinear systems theory is employed.
Lemma. Consider the autonomous system (3) with steady
state (xss , u ss ). Dene A as in (4). Then (xss , u ss ) is an
exponentially stable steady state of (3), if all eigenvalues of
A(xss , u ss ) lie in the open left half of the complex plane.
Conversely, the steady state is unstable, if A(xss , u ss ) has
at least one eigenvalue in the open right half of the complex
plane.
For a proof see Vidyasagar (1993), for example.
The above lemma helps to clarify the differences between
the two linearizations (4) and (7). The eigenvalues of the
linearization at a steady state (4) are those of A(xss , u ss ),
thus they are independent of u.
In contrast, the eigenvalues
of the linearization along a curve of steady states (7) are the
eigenvalues of A((u 1 ), u 1 ) which depend on u 1 . While the
linearization at a steady state (4) only allows to infer the
stability of the nonlinear system at the point of linearization
(and, by continuity, for values of u in a sufciently small
vicinity of u ss ), the linearization along a curve (7) can be
used to determine the stability of the nonlinear system for
nite ranges of u 1 U . This is illustrated with a simple
model describing a continuous stirred tank reactor.
Consider a model describing two sequential rst-order reactions P A, A B carried out in a stirred tank. P is
a pool chemical whose concentration is assumed to be constant. While the rate of the rst reaction is independent of
temperature, the second reaction obeys an Arrhenius temperature dependence, k(T ) exp(E/RT ). If stated in

stable
unstable
Hopf bifurcation

15


for all u 1 U . For simplicity of notation the constant


u 2,ss , ..., u nu ,ss are omitted from Eq. (5). This yields

4327

10

5
0
0

0.01

0.02

0.03

0.04

0.05


Fig. 1. Result of numerical paramerter continuation in .

dimensionless variables, the model reads





 =    exp
,
(1 + )



( a ).
 =  exp
1 + 

(8)

In Eq. (8),  is a dimensionless measure of the concentration


of A,  and a are the dimensionless temperature in the
reactor and the dimensionless temperature of the cooling
medium,  is the dimensionless concentration of P, and  is
dened by  = E/R Ta , where Ta is the temperature of the
cooling medium. As the model is used here for demonstrative
purposes only, the physics behind it are not discussed in
further detail. Instead, the reader is referred to Scott (1991)
for details.
Fig. 1 shows the result of a numerical parameter continuation of Eq. (8) with respect to u = . Given a known
 ,u
steady state (xss , u ss ), a close-by steady state (xss
ss ) for

u ss = u ss + u can be determined by applying Newtons
algorithm to 0 = f (x, u ss ) for xed u ss , using (xss , u ss )
as an initial guess. Loosely speaking, the initial guess
can be made as good as necessary by making u as
small as required. By repeating these so-called predictor
steps (xss , u ss ) (xss , u ss ) and Newton corrector steps
 ,u
ss ),a curve of steady states can be built
(xss , u ss ) (xss
up for the range  (0, 0.05), cf. Fig. 1. It is noted that
limit points at which a solution curve attains an extremum
in u can be handled if the curve  is parameterized by its
pseudo-arclength (Kuznetsov, 1998b).
The stability of the steady states shown in Fig. 1 has been
determined by evaluating the eigenvalues of the linearized
model and using the above Lemma as u =  was varied. The
result is shown in Fig. 2. As u is varied, the system can lose
stability, if a single real eigenvalue or a complex conjugate
pair of eigenvalues crosses the imaginary axis. These two
cases are related to saddle-node and Hopf bifurcations of
the system (Guckenheimer and Holmes, 1993; Kuznetsov,

4328

J. Hahn et al. / Chemical Engineering Science 59 (2004) 4325 4338


1
0.8
Hopf bifurcation

0.6
real parts

0.4
0.2
0
-0.2
-0.4
-0.6
-0.8
-1

0.01

0.02

0.03

0.04

0.05

Fig. 2. Real parts of the eigenvalues for the steady states shown in Fig.
1. Note that two real eigenvalues exist for  = 0. The two corresponding
branches merge at about  = 0.003 to form a single branch. For values of
 larger than roughly  = 0.003 a complex conjugate pair of eigenvalues
exists. The single branch marks the common real part of this complex
conjugate pair.

1
Hopf locus
0.8

a

0.6
0.4
unstable

stable

0.2
0
0

0.01

0.02

0.03

0.04

0.05

Fig. 3. Hopf locus in the (a , )-plane for model (8).

1998b). A discussion of the state space behavior that is related to these bifurcation points is beyond the scope of the
present paper. Here it is exploited that the stability boundary of the system can be analyzed by nding the loci of
Hopf and saddle-node bifurcations with respect to u.
Note
that eigenvalues were calculated for ease of illustration only.
In fact, a numerically expensive calculation of eigenvalues
can be avoided if test functions, which signal the crossing
of a stability boundary by a change in sign, are used instead
(Beyn et al., 2002).
Just as a curve of steady states was built up by parameter
continuation, curves of Hopf and saddle-node bifurcation
points can be traced. Fig. 3 shows the result of a continuation
of either of the Hopf points in Fig. 1. The curve in Fig. 3 was
obtained by applying parameter continuation to a standard

augmented Hopf system (Beyn et al., 2002). This gure is the


main result needed to illustrate how numerical continuation
and bifurcation analysis can, loosely speaking, provide more
than local information on stability. The curve of Hopf points
shown in Fig. 3 separates the region of the (, a )-plane
in which only stable steady states exist from a region in
which only unstable steady states exist. Note that the region
of stability is not only valid for any approximation to the
nonlinear model (8), but for the nonlinear model itself. This
is true even though the process of obtaining the information
visualized in Fig. 3 involved linearizations of model (8).
Further note that the information contained in Fig. 3 could
not have been obtained using a linearization of type (4),
since the eigenvalues of Eq. (4) are independent of u 1 = ,
u 2 = a . In contrast, the linearization along the curve of
steady states (7) captures the nonlinear dependency of the
stability boundary of model (3) on u 1 =  and u 2 = a .
The sequence of continuation, detection of saddle-node
and Hopf bifurcations, and continuation of saddle-node and
Hopf bifurcations can be repeated for critical points of higher
order, e.g. for a so-called cusp point found on a curve of
saddle-node bifurcations. Many types of higher order points
are related to exotic dynamic behavior such as chaos. While
these exotic points are less frequently of importance in engineering contexts, there are higher order critical points which
reveal information that can be exploited for practical purposes. In this work, sets of a particular type of degenerate
Hopf points bound regions in which no Hopf bifurcations
occur.
While both bifurcation analysis and nonlinear control theory deal with the stability of nonlinear dynamic systems,
the vast majority of the literature on bifurcation analysis
has nevertheless focused on open-loop processes. There are
several exceptions to this, one of them being investigations
of specic systems, where bifurcation analysis is performed
on the open-loop process in order to determine appropriate
operating points for designing a controller (for an example
see Zhang et al. (2002)). Another exception is the eld of
bifurcation control (Chen et al., 2000), however, the combination of elements from bifurcation analysis and control theory used in bifurcation control is considerably different from
the approach or even the goals of the technique presented in
this paper. In bifurcation control the predominant theme is
the control of the dynamic behavior in the parametric vicinity of a bifurcation. This bifurcation is a critical point of the
open-loop system, and the control is added to achieve a desired closed-loop behavior whenever the controlled system
enters the parametric vicinity of the open-loop critical point.
A typical example is the transformation of a subcritical into
a supercritical Hopf bifurcation in aircraft control (Chen
et al., 2000). Despite addressing the quite special case of operating a system in the vicinity of a critical point, bifurcation
control has been applied to an industrial chemical engineering process (Recke et al., 2000). Recent work in bifurcation
control addresses the deliberate introduction of bifurcation
points and the exploitation of the accompanying exchange

J. Hahn et al. / Chemical Engineering Science 59 (2004) 4325 4338

of stability for stabilization (Mnnigmann and Marquardt,


2002a). Another type of investigation which performs bifurcation analysis on systems containing controllers deals with
specic processes where stability analysis is performed at
one operating point, but for the case of variations in a parameter of a model (Paladino and Ratto, 2000) or for the gain
of a PI controller (Giona and Paladino, 1994) or the gain of
a proportional-only controller (Cibrario and Lvine, 1991).
Along the same lines, it has been noted in Ananthkrishnan
et al. (2003) that for an axial compressor, there may exist
a threshold value for certain controller tuning parameters
in order to eliminate the occurrence of Hopf bifurcations.
The types of bifurcations that can occur in closed-loop systems under PID control have been investigated by Chang
and Chen (1984).
In the analysis of open-loop systems, bifurcation analysis
and continuation have been used to determine which regions
of the process parameter space allow for stable and robust
process operation in the way sketched above. The present
paper extends this idea to closed-loop systems. By varying
both, open-loop process and controller/observer parameters
in continuation, bifurcation analysis is used to determine
the regions of the closed-loop system parameter space in
which the desired, i.e. stable and robust, dynamic behavior
exists. Additionally, by using the set points of the system as

u=

Lrf h(x)

4329

given by

 1 = 2 ,
 2 = 3 ,
..
.
 r = Lrf h(x) + Lg Lr1
f h(x)u,
 = q(, ),
y = 1 .

(11)

The map between the input and the output can be linearized
by choosing a static state feedback control law
u=

v Lrf h(x)

(12)

Lg Lr1
f h(x)

such that the rth equation of (11) becomes  r = v. It is possible to place the poles of the closed-loop transfer function
for the linearized subsystem  dened by Eq. (11) in the
complex plane by choosing an appropriate feedback v. However, in most cases, the precise location of the closed-loop
poles is less important than the distance of the poles from
the imaginary axis. Therefore, only one tuning parameter 
which represents the time constant of the closed-loop system, is used to shape the closed-loop response in this paper.
A feedback linearizing controller in terms of the original
states is then given by

r
( r1
)

(r )

r1
1
1
 Lf h(x) r1 Lf h(x) + r (ysp h(x))
.
Lg Lr1
f h(x)

(13)

bifurcation parameters, it is possible to determine regions


of stability not just for the parameter space but for the state
space corresponding to the operating region of a process.

When this control law is applied to the process, the closedloop transfer function between the system output y and the
set point ysp becomes

2.2. Feedback linearization

y(s)
1
=
ysp (s) (s + 1)r

(14)

Two main categories of designing controllers via feedback linearization can be identied: inputoutput linearization and state-space linearization. The presentation in this
paper will exclusively focus on the former because it is more
generally applicable and will result in a linear inputoutput
behavior of the system if no model mismatch is present.
Consider a SISO control-afne nonlinear system with n
states of the form

under the assumption that the initial condition is y(0) =


ysp (0).
It is also possible to include integral action in the controller in order to compensate for possible inaccuracies in
the model. The feedback linearizing controller with integral
action is given by

x = f (x) + g(x)u,
y = h(x).

u=
(9)

If this system has a well-dened relative degree r, then it


can be transformed into normal form via a diffeomorphism
[T , T ]T = (x). The  coordinates are dened as

k = k (x) = Lk1
f h(x),

1k r

(10)

and the j = r+j (x), 1j n r correspond to the internal dynamics of the process (Isidori, 1995; Kravaris
and Kantor, 1990). The normal form of the system is then

( r+1
r )

( r+1 )

2
Lr1
f h(x) r1 Lf h(x)
Lg Lr1
f h(x)
r+1
t
( 1 )
1
r (ysp h(x)) + r+1 0 (ysp h(x)) d

Lrf h(x)

Lg Lr1
f h(x)

(15)

resulting in the closed-loop transfer function


y(s)
(r + 1)s + 1
=
ysp (s)
(s + 1)r+1
under the condition that y(0) = ysp (0).

(16)

4330

J. Hahn et al. / Chemical Engineering Science 59 (2004) 4325 4338

It is important to note that the transfer functions shown


in Eqs. (14) and (16) only represent the closed-loop system
behavior if no model mismatch exists between the plant
and the model used for controller design, or if the model
mismatch satises the matching conditions up to a degree
of at least r (Henson and Seborg, 1997). In this case, the
closed-loop transfer functions are linear and have all their
poles at 1 . Therefore, it is possible to guarantee nominal
performance with these types of controllers if the internal
dynamics of the process, given by the subsystem denoted by
 in Eq. (11), are stable. It should also be pointed out that the
resulting closed-loop transfer functions shown in Eqs. (14)
and (16) are identical to the lters used for linear internal
model control (Morari and Zariou, 1989), which in turn
are equal to the closed-loop transfer functions for minimum
phase processes.
Since this investigation specically focuses on the effect
of model uncertainty on controller design and tuning, the
controller implementation shown in Eq. (15) will be used for
the feedback linearizing controller. Due to model mismatch,
it will not be possible to exactly achieve the closed-loop
response shown in Eq. (16). However, the integrating term
in the controller will ensure that the desired set point can be
reached.
For the implementation of feedback linearizing controllers
it is often postulated that the internal dynamics of the process are stable and that the values of the states are known
exactly (Hahn et al., 2003, 2004). The validity of the former
assumption can easily be analyzed as part of the proposed
method while the latter assumption can be relaxed by designing an observer which estimates the states of the system
and provides their values to the controller.
Other methods for designing controllers for nonlinear system under the inuence of model uncertainty have been proposed in the literature (see e.g. Arkun and Calvet (1992) or
El-Farra and Christodes (2001) for specic techniques as
well as Marino and Tomei (1995) or Henson and Seborg
(1997) for a review of several methods). These techniques
use different approaches to determine if tuning parameters
can be found for robust control of nonlinear processes as
well as conditions which tuning parameters have to satisfy
for achieving robust stability. However, there is a major difference in the motivation behind these approaches and the
one presented in this work. The bifurcation analysis-based
procedure does not just result in conditions for achieving
stable operations, but it performs an explicit computation of
the values for the controller tuning parameters resulting in
stable operating points.

and an observer
x)
x = f(x)
+ g(
x)u
+ K(h(
y),

y = h(x),

2.3. Observer design

x = f(x,
p)
+ g(
x,
p)u
+ K(x,
p,
p)(
y y),

y = h(x,
p),

u = k(x,
p,
, ysp )

The assumption that the states of system (9) are known is


commonly made when a state feedback linearizing controller
as given by Eq. (13) or (15) is designed. However, the values
for all the states are usually not known for a real process

(17)

needs to be implemented which estimates the values of the


states (Kazantzis et al., 2000). The functions f, g,
and h
from Eq. (17) correspond to estimates of the functions f, g,
and h of plant (9). The nonlinear controller
u(x,
, ysp )
=

x)
Lr h(

f

( r+1
r )

( r+1 )

2
x)
x)
h(
r1
Lr1
Lf h(

f

Lg Lr1
h(x)

( r+1
1 )

r

t
1
x))+
x))
(ysp (0)h(

(y h(
d
r+1 0 sp
,
r1
Lg L h(x)

(18)

will then use the values of the estimated states x.


This type of
controller in combination with an observer is called an output feedback linearizing controller (Mahmoud and Khalil,
2002). It should be noted that the separation principle which
allows to independently design the controller and the observer in order to achieve stability for linear systems does
generally not hold for nonlinear systems. Therefore, observer
and controller design and tuning have to be performed simultaneously for nonlinear processes.
The choice of a procedure for computing K determines
the type of nonlinear observer that is designed.
3. Bifurcation analysis of systems under output
feedback linearizing control
This section builds upon the reviewed material and explains how bifurcation analysis can be interpreted in the context of robustness analysis of controlled nonlinear systems.
In order to proceed, a notation that enables presenting
bifurcation analysis and controller/observer design within
the same context is introduced.
3.1. Application to systems with observer and controller
Consider an open-loop nonlinear dynamic system of the
form
x = f (x, p) + g(x, p)u,
y = h(x, p),
(19)
where x, y, u, and p represent the states, the outputs, the
inputs, and the parameters of the system, respectively.
It is possible to design an observer and a controller

(20)

for the system in order to stabilize the process or to impose


desired performance characteristics. The static control law

J. Hahn et al. / Chemical Engineering Science 59 (2004) 4325 4338

can be a function of the estimate of the systems states x and


of the process parameters p.
Additionally, the control law
introduces controller tuning parameters  and new inputs in
the form of set points ysp . Note, that the observer gain K
can be a function of the estimated position in state space x
as well as of the estimates of the plant parameters p and of
observer parameters p.

The resulting closed-loop system is then given by


x = f (x, p) + g(x, p)k(x,
p,
, ysp ),
f(x,
x=
p)
+ g(
x,
p)k(
x,
p,
, ysp )

+ K(x,
p,
p)(
h(x,
p)
y),
y = h(x, p).

(21)

The parameters of the system, of the controller, and of the


observer as well as the inputs to the controller can be concatenated into a new vector p = (p, p,
p,
, ysp ) and the
states of the plant as well as of the observer can be lumped
into one vector x = (x, x).
Using this substitution results in
the system
x = f(x,
p),

x,
y = h(
p).

(22)

The ODEs of Eq. (22) have the same form as the equations
to which bifurcation analysis is typically applied.
Summarizing, there are several steps involved in performing bifurcation analysis of controlled nonlinear systems:
Identify parameters p and inputs u for the open-loop process.
Choose a controller structure for the system. Additionally,
design an observer if the controller requires information
about states which cannot be measured. The output of the
controller serves as input to the system and has to be removed from the set of variables to be used for bifurcation
analysis. Instead, the inputs to the controller as well as
controller and observer tuning parameters are added as
additional variables.
Perform bifurcation analysis using parameters of the process, set points as well as controller and observer parameters.
It is important to note that the set point of the system is one of
the bifurcation parameters. Since the bifurcation parameters
can be varied over nite ranges in parameter continuation,
this implies that the analysis can be performed over an entire
operating region of the process rather than for a particular
xed value of the set point.
However, it should be pointed out that it is assumed for
this investigation that the set point is not changed with a
high frequency. This assumption results from the bifurcation
analysis which returns information about equilibrium points
of the system but not about stability under the inuence of
time-varying excitations. For a discussion of the problem
class that can be addressed with bifurcation theory-based

4331

approaches the reader is referred to (Mnnigmann and Marquardt, 2003).


3.2. Interpretation of results from robustness analysis
The purpose of performing bifurcation analysis on controlled nonlinear systems is to determine robustness properties of a chosen controller/observer conguration in the presence of parametric uncertainty and unmodeled dynamics.
The analysis can return Hopf and limit point curves that separate regions with stable steady states from unstable regions
in the parameter space. It is often possible to determine that
a system will remain stable over the entire operating region
(represented by the set points of the system) for a certain
amount of plant-model mismatch if the controller/observer
tuning parameters are chosen within a specic range. If such
a situation is encountered, then it is possible to state that the
dynamic system will be stable for the investigated scenario,
since no regions of instability are present. For reasons of
robustness, it is preferable to keep a certain distance in the
parameter space between the region of instability and the
desired set of tuning parameters. Therefore, it is possible
to compare different controller/observer congurations with
regard to robustness of the closed-loop system by evaluating the size of the region in parameter space for which the
system has only stable steady states.
However, it needs to be pointed out that the magnitude of
the region of stability is only one possible measure for robustness of a nonlinear system. Investigation of other properties of controlled nonlinear systems, e.g. the analysis of
the behavior of a process subject to time-varying disturbances, will also result in important information. The bifurcation analysis-based technique is meant as a tool which
complements existing methods (Mahmoud and Khalil, 2002;
Marino and Tomei, 1995; Christodes, 2000; Arkun and
Calvet, 1992) by explicitly computing bounds which certain
types of model uncertainty impose on a control system.
3.3. Bifurcation analysis-based controller tuning
Since the proposed technique will return stability boundaries for the controller and observer tuning parameters over
the operating region of a process, this information can be
used in order to tune the controller/observer combination
while ensuring that all steady states within this operating region will be stable. The following steps should be included
in the controller/observer tuning process:
(1) Design and implement an observer/controller combination on the simulated process.
(2) Analyze the internal dynamics of the closed-loop system
using bifurcation analysis. This should be performed at
the operating point as well as over the entire operating
region in order to determine if the internal dynamics will
remain stable over this region.

4332

J. Hahn et al. / Chemical Engineering Science 59 (2004) 4325 4338

(3) If the controller is a feedback linearizing controller then


it can be tuned to satisfy a nominal performance requirement.
(4) Identify the main sources of parametric uncertainty in
the model. Subsequently, analyze the closed-loop system using bifurcation analysis under the assumption of
parametric uncertainty. This can result in restrictions on
the controller tuning parameters.
(5) Perform bifurcation analysis on the closed-loop system
under the assumption of unmodeled dynamics. This investigation might place further bounds on the controller
tuning parameters.
(6) Investigate the region for the controller tuning parameters for which the closed-loop system remains stable
under the worst possible combination of parametric uncertainty and unmodeled dynamics over the entire operating region. Requiring stability of the closed-loop system over the entire operating region is important because
bifurcation analysis only results in steady state information, and it has to be ensured that the system trajectories cannot leave the region of attraction of the steady
state operating point. Therefore, this requirement is an
important point in order to guarantee robust stability of
the controller.
(7) If the controller/observer tuning parameters that satisfy
a nominal performance requirement can also guarantee
stability over the entire operating region, then they can
be kept. Otherwise, the controller/observer has to be retuned in order to ensure stable operations. It should be
pointed out that it is desirable to use controller/observer
tuning parameters that do not lie close to a region of
instability of the closed-loop process in order to also
achieve good performance even under the inuence of
plant-model mismatch. If there is a gap of several orders of magnitude between the smallest and the largest
value of the tuning parameter then it is recommended
to stay closer to the lower value in order to achieve a
faster response. Ultimately, the dynamics of the openloop process has also to be considered for tuning the
controller/observer.
Following these steps will result in a controller/observer
combination that will result in stable operating points over
the entire region of operation.
It is stressed that Step 6 of the above procedure corresponds to an analysis of cusp and degenerate Hopf points
involving more than two parameters. Repeated calculation
of curves in two parameters will become tedious if extremal
points must be found w.r.t. more than three parameters. It
is worth noting that design methods based on bifurcation
theory (Mnnigmann and Marquardt, 2000b) are available
which can deal with a larger number of parameters than bifurcation analysis. However, this does not replace the task
of appropriately parameterizing the system, the controller,
and the observer since the computational effort grows with
the number of parameters that are allowed to vary.

4. Investigation of three controllers


In this section the bifurcation analysis-based method is
applied to three different controllers which operate on a
continuously stirred tank reactor:
a state feedback linearizing controller. No observer is required for this type of controller, however, the assumption
of knowing the values of both states is not very realistic
in practice;
an output feedback linearizing controller where the observer gain is chosen based upon the nominal operating
point and is kept constant;
an output feedback linearizing controller where the observer gain is varied depending upon the location in state
space (Zeitz, 1987).
The results for the state feedback linearizing controller are
compared to the restrictions determined for the controller
congurations containing observers. It will be shown for this
example that model uncertainty has a signicantly stronger
effect on the observer/controller conguration than on the
controller alone. These results can serve as an indicator that
investigations of advanced control strategies should take into
account all of the required components of a control system
and not make unrealistic assumptions about the availability
of state information.
4.1. Model description
Consider a continuous stirred tank reactor (CSTR) for
an exothermic, irreversible reaction, A B (Uppal et al.,
1974). Assuming constant liquid volume, the following dynamic model can be derived based upon a component and
an energy balance



q 
E

CA =
CAf CA k0 exp
CA ,
(23)
V
RT


 H
E
q 

k0 exp
Tf T
CA
T=
V

C p
RT
UA
+
(Tc T ).
(24)
V
C p
The values of the parameters and the nominal operating conditions for this process are shown in Table 1. The temperature of the cooling uid, Tc , can be manipulated and the
reactor temperature, T, is measured. This results in a system
consisting of two states with a single input and a single output. While manipulation of Tc may not represent the most
realistic scenario for this process, it is nevertheless chosen
since this model has been widely used for controller design
including several publications on feedback linearizing controllers (Henson and Seborg, 1997). The bifurcation diagram
of the open-loop system is shown in Fig. 4a. The equilibria
of the system consist of two stable branches and one unstable branch connecting the two stable ones. The system

J. Hahn et al. / Chemical Engineering Science 59 (2004) 4325 4338


Table 1
Parameters for the CSTR

400
390

Variable

Value

Variable

Value

L
100 min

E
R

8750 K

CAf

1 mol
L

k0

1
7.2 1010 min

Tf

350 K

UA

100 L

Tc

5 104 minJ K
300 K

g
1000 L

CA

0.5 mol
L

340

Cp

J
0.239 gK
J
5 104 mol

350 K

330

H

4333

Stable Branch
Unstable Branch
Limit Point
Hopf Point

380
370

T [K]

360
350

320
310
300
270

4.2.1. State feedback linearizing controller


When a state feedback linearizing controller like the one
given by Eq. (15) is designed for this process it results in
the following feedback control law

330

310

320

330

370
360
350
340
330
320
310
300
280

290

300
Tc [K]

Fig. 4. Bifurcation diagram of the nominal open-loop (a) and closed-loop


(b) system. The dashed lines correspond to the nominal operating point
from Table 1.

UA
V
Cp

(26)

320

380


E
2
1 t
Vq (Tf T ) +
CHp k0 exp( RT
)CA + VU
A
Cp T +  (Tsp T ) + 2 0 (Tsp T ) d

Tc = u.

310

390

(b)

4.2. Controller congurations

300
Tc [K]

270

u=

290

400

Tsp = T [K]

has two limits points and one Hopf point. The nominal operating point dened by the parameters in Table 1 lies on
the open-loop unstable branch. When a feedback linearizing controller (either with or without observer) is designed
for this process then the unstable branch of the equilibria
is stabilized. The results for the closed-loop case and no
model uncertainty are shown in Fig. 4b. Note, that in Fig. 4
the temperature of the cooling uid Tc serves as the bifurcation parameter for the open-loop case while the set point
Tsp is varied for the closed-loop system. Since the set point
of the process is equal to the temperature of the system for
the closed-loop case, it is possible to make statements about
the stability of the system in a region of the state space by
performing bifurcation analysis in the parameter space.

280

(a)

(25)

the states

Assuming that there is no mismatch between the plant and


the model, the controller of the form of Eq. (25) results in
a system that has a stable inputoutput behavior as well as
stable internal dynamics for any value of  and any set point,
Tsp , within the operating region. This can easily be veried
by applying bifurcation analysis to the closed-loop system.
The results of this analysis are shown in Fig. 4b.



q
E

C A = (CAf CA ) k0 exp
V
R T
C A + K1 (T T ),


q
H
E
T = (Tf T )
C A
k0 exp
V

C p
R T
UA
+
(Tc T ) + K2 (T T ).
V
C p

4.2.2. Output feedback linearizing controller with


nonlinear observer
If not all states of a process can be measured, then an
observer needs to be designed which estimates the values of

A feedback linearizing controller can then make use of the


values of the estimated states. Hence, the estimated quantities C A and T replace the true states CA and T in Eq. (25)
resulting in an output feedback linearizing controller.

(27)

(28)

4334

J. Hahn et al. / Chemical Engineering Science 59 (2004) 4325 4338

The locations of the eigenvalues of the observer are computed from

1 =

observer

2 =

1.001

observer

(29)

where observer refers to the desired time constant of the


observer. Note that the eigenvalues 1 and 2 are distinct
but the numerical values are similar since this approach does
not allow for repeated eigenvalues.
It is possible to write an explicit expression for computing
the observer gain for a two state process (Zeitz, 1987)
K1 (x, u, 1 , 2 )


*f1
*f1

1 2
1
2
*f1
*x1
*x1
=
+

*
f
2
*x 2

*x1

(30)

x,u

*f1 *f2
+
K2 (x, u, 1 , 2 ) = ( 1 + 2 ) +
*x 1
* x2

(31)

x,u

where f1 and f2 correspond to the model of the right hand


side of Eqs. (23) and (24), respectively, and x1 = C A , and
x2 = T . Eqs. (30) and (31) can be evaluated at the nominal
operating point (x0 , u0 ) and for i , i= 1, 2, given by Eq.
(29) resulting in observer gains which are only a function
of observer . However, i , i= 1, 2, from Eq. (29) will only
correspond to the location of the eigenvalues of the observer
at this nominal operating point. For the rest of the paper
this type of controller will be referred to as output feedback
controller A.
Another observer design strategy is to re-evaluate Eqs.
(30) and (31) at C A and T each time the observer subroutine
is called. Thereby, it is possible to have constant values for
the eigenvalues of the observer, given by Eq. (29) and independent of the process operating conditions. However, this
comes at the cost of variable observer gains K which have
to be recomputed at each time step. This type of observer is
called an extended Luenberger observer (Zeitz, 1987). The
feedback linearizing controller in conjunction with an extended Luenberger observer will be referred to as output
feedback controller B for the remainder of this work.
The bifurcation diagram shown in Fig. 4b results if either one of the output feedback controllers is applied to the
system.
4.3. Investigated model uncertainty
4.3.1. Parametric uncertainty
Every process model contains model parameters whose
values can only be known up to a certain degree. Therefore, it is safe to say that parametric uncertainty exist for
any model. However, the effect that parametric uncertainty
has on controller and observer design and tuning does depend upon which parameters have the largest uncertainty,
the magnitude of the uncertainty, as well as the type of controller/observer which is used.

For this investigation it will be assumed that the heat transfer coefcient UA varies by up to 10% around its nominal
value. Uncertainties in other parameters have also been investigated for the state feedback linearizing controller. They
result in similar or lesser restrictions for the controller than
the investigated case. In order to be able to present the results
in a clear and concise manner, only parametric uncertainty
of UA is considered in this work.
4.3.2. Unmodeled dynamics
Another form of model mismatch that needs to be investigated is unmodeled dynamics. Since no model can describe
every detail of a process with perfect precision, it is commonly assumed that the fast dynamics of the process can be
neglected. While this is generally a good assumption, it does
lead to a bound on the achievable closed-loop performance.
In order to investigate these restrictions, bifurcation analysis
can be performed on a system that contains the most important part of the fast dynamics while a controller that has
no knowledge about this dynamics is used to control it. For
this work, the plant model is augmented by the following
two equations which describe the actuator dynamics as an
overdamped second-order process

v Tc = Tc + z
v z = z + u

(32)

where v corresponds to the time constant of the cooling


system. The equations in (32) replace the original Eq. (26)
for the following investigation. Unmodeled dynamics with
a time constant of up to 0.02 min have been investigated.
The goal is to determine the restrictions imposed upon controller/observer tuning by this form of model uncertainty.
4.4. Comparison of the controllers
In order to perform a comparison of the robustness properties of the different controllers applied to the CSTR, the
controllers have to be implemented on the model which includes the uncertainty. In a next step it is checked via bifurcation analysis if the closed-loop system is stable for a
chosen set of parameters. The controller or observer tuning
parameter is then varied until a stability boundary, which in
this case study was given by a Hopf point, is determined.
Starting from the Hopf point, it is possible to compute a
Hopf curve by varying the set point as an additional bifurcation parameter. The resulting Hopf curve separates a region in which only stable steady states exist from a region
in which all steady states are unstable. This step highlights
one of the strong points of the proposed methodology: while
many techniques can compute a point where the closed-loop
system becomes unstable, it is just as simple to compute an
entire curve of critical points using bifurcation analysis on
a controlled nonlinear system. The results for a state feedback linearizing controller are shown in Fig. 5 in which
two Hopf curves are depicted: the Hopf curve for low values of  results from unmodeled dynamics in the process

J. Hahn et al. / Chemical Engineering Science 59 (2004) 4325 4338


440

440

420

420
400

400
380

Stable Region
380

Unstable Region

360

360
Unstable Region

340

340
320

Tsp= T [K]

Tsp= T [K]

4335

320

Stable Region

300

300
280
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
[min]

5.5

6.5

7.5

8.5

9.5

280
10

[min]

Fig. 5. Regions of stability based upon the controller tuning parameter .

(v = 0.02 min) and the curve on the right is produced by


parametric uncertainty of +10% in the model parameter UA,
which corresponds to the strongest limitation imposed by
uncertainty in this model parameter. It is important to point
out that unmodeled dynamics results in a bound representing the most aggressive controller that can be implemented
while uncertainty in the model parameters results in an upper bound for the tuning parameters of feedback linearizing controllers. While this may seem counterintuitive, it can
be explained by the control law structure that is unique to
feedback linearizing controllers. A detailed explanation of
this phenomenon is provided in the Appendix. Once stability boundaries have been determined, it needs to be investigated how these bounds vary with a change in the magnitude of the model uncertainty. It was found that if the model
uncertainty is less than was assumed for the investigation
then the bounds are relaxed and the process will remain stable. The value for the tuning parameter  can be between
0.059 min <  < 7.2 min for the state feedback linearizing
controller and result in stable operation.
For the output feedback linearizing controllers, the observer tuning parameter observer has to be varied in addition
to the controller tuning parameter . This will result in a set
of curves as shown in Fig. 6 for the case of parametric uncertainty of +10% in the model parameter UA for the output feedback controllers. The gure showing the bounds for
10% uncertainty in the model parameter has been omitted since this will result in lesser restrictions on controller
tuning.
Similarly, the Hopf curves for the closed-loop process
for unmodeled dynamics with v = 0.02 min are shown in
Fig. 7. Several conclusions can be drawn from this analysis:
The observer tuning parameter has a stronger inuence
on the stability of the system than the controller tuning
parameter.
The more aggressive the controller, the less restrictive
the bounds on the observer for parametric uncertainty.

For unmodeled dynamics the effect of the aggressiveness


of the controller on the observer parameter bound is the
opposite.
However, it has to be kept in mind that both observer and
controller tuning parameters should be in an appropriate
range when compared to one another as well as in comparison to the dynamics of the open-loop process.
Since it is not easily possible to compare results if several parameters are varied, the controller tuning parameter
will be xed at  = 0.5 min for the remainder of the investigation. This value was chosen based upon the dynamics of
the open-loop process as well as on the restrictions for the
state feedback linearizing controller.
The bounds for the observer tuning parameter for both
output feedback controller A and B are given in Table 2.
From the values shown in the table it can be concluded that
the bounds on observer tuning are much tighter than for controller tuning. While  for the state feedback linearizing controller could vary over more than two orders of magnitude
and still result in stable operation, the region for observer
for the output feedback controller A only allows the tuning
parameter to vary by a factor of less than ve. For the output feedback controller B the region is larger than for controller A, but the bound imposed by parametric uncertainty
is even more stringent. Since it is not uncommon in practice
to design the observer about twice as fast as the controller,
this upper bound on observer presents a severe limitation for
 = 0.5 min. At the same time, the lower bound of 0.011 min
is not a very realistic scenario since this results in an observer
which is faster than the unmodeled dynamics of the process.
An investigation for  = 0.25 min has also been performed
which returned similar results.
The following conclusions can be drawn for this case
study:
model uncertainty results in signicantly tighter bounds
for the output feedback linearizing controllers than for the
state feedback linearizing controller.

4336

J. Hahn et al. / Chemical Engineering Science 59 (2004) 4325 4338


400

400
= 0.25 [min]
= 0.5 [min]
= 1 [min]
= 5 [min]

390

380

370

Tsp = Tobserver [K]

Tsp = Tobserver [K]

380

= 0.1 [min]
= 0.25 [min]
= 0.5 [min]
= 1.0 [min]

Stable Region

390

360
Unstable Region

350
340
330

370
360
350
340
330
ta
e

bl

310

ns

320

310

320

eg

R
io
n

300

300
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

observer [min]

observer [min]


Fig. 6. Hopf curves for output feedback linearizing controllers and parametric uncertainty (U
A = 1.1 UA). The results for output feedback controller A
are shown on the left and for output feedback controller B on the right.

400

400
= 0.1 [min]
= 0.25 [min]
= 0.5 [min]
= 1.0 [min]

380

380

370

370

360
= 0.25 [min]
= 0.5 [min]
= 1 [min]
= 5 [min]

350
340
330

Unstable Region

390

Unstable Region

Tsp = T [K]

Tsp = T [K]

390

360
350
340
330

320

Stable Region

320

Stable Region

310

310

300

300
0

0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2

0.01

0.02

0.03

0.04

0.05

0.06

observer [min]

observer [min]

Fig. 7. Hopf curves for output feedback linearizing controllers and unmodeled dynamics (v = 0.02 min). The results for output feedback controller A
are shown on the left and for output feedback controller B on the right.

Table 2
Bounds on observer tuning parameter observer

Nonlinear observer (27)(31)


Extended Luenberger observer

Unmodeled
dynamics (min)

Parametric
uncertainty (min)

0.094
0.011

0.42
0.21

observer tuning has a stronger effect on stability of the


system than controller tuning for the output feedback controllers.
while uncertainty in the model parameters has only a minor effect on stability for the controller, it has a signicant
effect on the observer/controller combination.
designing a controller which uses more knowledge about
the process, like output feedback controller B, does

not necessarily result in better performance/robustness


over simpler controllers when model uncertainty is
present.

5. Conclusions
This paper presented a method for analyzing robustness of
controlled nonlinear systems. The technique is based upon
applying bifurcation analysis to the closed-loop system. The
controller and observer are designed using a nominal model
while mismatch can exist between the nominal model and the
plant. Bifurcation analysis is then used to determine limits
for the controller and observer tuning parameters when the
controller is acting on the plant.
This technique has been illustrated with a case study
where a state feedback linearizing controller was compared

J. Hahn et al. / Chemical Engineering Science 59 (2004) 4325 4338

to two output feedback linearizing controllers with different


levels of complexity for the observer design. The same types
and magnitudes of model uncertainty were investigated for
all three controllers. It was found that the plant-model mismatch resulted in bounds which were not tight for the state
feedback linearizing controller. At the same time the uncertainty had a more profound impact on the two output feedback linearizing controllers as can be concluded from the
stringent conditions placed on the observer tuning parameter. In fact, these bounds were so tight, that it is questionable
if the output feedback linearizing controllers can be used
in practice for this process because additional model uncertainty may also be present. This analysis highlights that
the often employed assumption for controller design that all
states of a process are known is not always a good one. At
least for the investigated case it can be stated that model uncertainty has a stronger effect on the observer tuning than
on controller tuning.
It is worth pointing out that this analysis technique does
not make assumptions about stability of the model to be
controlled, nor is it restricted to specic types of controllers.
At the same time it should be noted that a requirement for
this method is that the uncertainty can be parameterized and
that disturbances to the system are assumed to be constant
or varying very slowly for the analysis.

Appendix A.
It has been pointed out in the paper that choosing a large
value for the controller tuning parameter  will have a negative effect on stability of the closed-loop system in the presence of parametric uncertainty. The mathematical explanation for this behavior is given below.
Consider a system with n states and relative degree r
which can be transformed into a normal form given by

 1 = 2 ,
 2 = 3 ,
..
.
 r = Lrf h(x) + Lg Lr1
f h(x)u,
 = q(, ),
y = 1 .

Lrf h(x)

Lg Lr1
f h(x)

v
Lg Lr1
f h(x)

resents the (possibly nonlinear) internal dynamics that are


unobservable:


1
0 1 0 ... 0
0
1

2 0 0 1 ... 0 2 0

. . . . .
. = .. .. ..
. . ... .. + ... v,
(35)
.
.
 0 0 0 ... 1 0

r1
r1
1
0 0 0 ... 0
r
r

 = q(, ),
y = 1 .
If v is chosen to be equal to zero, as it will be the case when
the closed-loop time constant  in Eq. (13) approaches innity, then this will result in a system where all the eigenvalues of the subsystem 1 ,...,r are identical to zero. Such
a feedback law will drive the system to the stability limit.
If there exists plant-model mismatch, then the relationship
between u and y will likely not be exactly linearized and
some of the eigenvalues of the closed-loop system may lie
in the open right-half of the complex plane. Therefore, the
feedback law shown in Eq. (34) needs to have an expression
for v that places the eigenvalues of the subsystem 1 ,...,r
in the open left-half complex plane and at a sufcient distance from the imaginary axis. This can be achieved by a
feedback law of the form shown in Eq. (13) with a carefully
determined value of :
 0
1
0
...
0 1
1
0
1
...
0 2
2 0
. ..
..
..
.. .
.
.
. = .

.
.
.
.
.
..
 0

0
0
...
1
r1
r1
r
r
r
( r1 )
(
(
)
)
1
1
2
r
r
r r1 r2 ... 

0
0
.
.
+
(36)
. ysp ,
0
1
r

 = q(, ),
y = 1 .

(33)

The map between the input and the output can then be linearized by choosing the static state feedback control law (12)
u(t) =

4337

(34)

such that the rth equation of (33) becomes  r = v. This feedback law consists of two parts: the rst term eliminates the
dynamic behavior of the nonlinear system while the second
term imposes the desired closed-loop behavior. This results
in the following linear inputoutput system, where  rep-

It is important to observe for feedback linearizing controllers


that choosing a closed-loop time constant  that approaches
innity will not result in a system with good robustness
properties. At the same time, a very small value of  will
place the eigenvalues far from the imaginary axis, also resulting in a loss of robustness. While the second observation
is common knowledge for designing linear controllers, the
effect of high values of  on the robustness of the closedloop system is specic to feedback linearizing controllers.
References
Ananthkrishnan, N., Vaidya, U.G., Walimbe, V.W., 2003. Global stability
and control analysis of axial compressor stall and surge phenomena
using bifurcation methods. Proceedings of the Institution of Mechanical
Engineers. Part A- Journal of Power and Energy 217, 279286.

4338

J. Hahn et al. / Chemical Engineering Science 59 (2004) 4325 4338

Arkun, Y., Calvet, J.-P., 1992. Robust stabilization of input/output


linearizable systems under uncertainty and disturbances. AIChE Journal
38 (8), 11451156.
Bequette, B.W., 1991. Nonlinear control of chemical processes: a review.
Industrial and Engineering Chemistry Research 30, 13911413.
Beyn, W.-J., Champneys, A., Doedel, E., Govaerts, W., Kuznetsov, Yu.A.,
Sanstede, B., 2002. Numerical continuation, and computation of normal
forms. Fiedler, B. (Ed.), Handbook of Dynamical Systems, Vol. 2.
Elsevier Science, North-Holland, pp. 149219.
Chang, H.-C., Chen, L.-H., 1984. Bifurcation characteristics of nonlinear
systems under conventional PID control. Chemical Engineering Science
39, 11271142.
Chen, G., Moiola, J.L., Wang, H.O., 2000. Bifurcation control: theories,
methods, and applications. International Journal of Bifurcation Chaos
10 (3), 511548.
Christodes, P.D., 2000. Robust output feedback control of nonlinear
singularly perturbed systems. Automatica 36, 4552.
Cibrario, M., Lvine, J., 1991. Saddlenode bifurcation control
with application to thermal runaway of continuous stirred tank
reactors. In: Proceedings of the 30th Control Decision Conference
pp. 15511552.
Doedel, E.J., Paffenroth, R.C., Champneys, A.R., Fairgrieve, T.F.,
Kuznetsov, Yu.A., Sandstede, B., Wang, X., 2001. AUTO 2000:
continuation and bifurcation software for ordinary differential equations
(with HomCont). Technical Report, California Institute of Technology.
El-Farra, N.H., Christodes, P.D., 2001. Integrating robustness optimiality
and constraints in control of nonlinear processes. Chemical Engineering
Science 56, 18411868.
Giona, M., Paladino, O., 1994. Bifurcation-analysis and stability of
a controlled CSTR. Computers and Chemical Engineering 18,
877887.
Guckenheimer, J., Holmes, P., 1993. Nonlinear Oscillations, Dynamical
Systems, and Bifurcations of Vector Fields. Springer, New York.
Hahn, J., Mnnigmann, M., Marquardt, W., 2003. Robust tuning
of feedback linearizing controllers via bifurcation analysis. In:
Proceedings of the ADCHEM 2003. pp. 525530.
Hahn, J., Mnnigmann, M., Marquardt, W., 2004. On the use of bifurcation
analysis for robust controller tuning for nonlinear systems. Journal of
Process Control, in press.
Henson, M.A., Seborg, D.E., 1991. Critique of exact linearization
strategies for process control. Journal of Process Control 1,
122139.
Henson, M.A., Seborg, D.E., 1997. Feedback linearizing control. In:
Henson, M.A., Seborg, D.E. (Eds.), Nonlinear Process Control. Prentice
Hall, Upper Saddle River, NJ.
Isidori, A., 1995. Nonlinear Control Systems. 2nd Edition. Springer, New
York.
Kazantzis, N., Kravaris, C., Wright, R.A., 2000. Nonlinear observer design
for process monitoring. Industrial Engineering Chemistry Research 39,
408419.

Kravaris, C., Kantor, J.C., 1990. Geometric methods for nonlinear process
control. 2. Controller synthesis. Industrial Engineering Chemistry
Research 29, 23102323.
Kuznetsov, Yu.A. 1998a. CONTENTIntegrated environment for
analysis of dynamical systems. Tutorial. Dynamical Systems
Laboratory, Centrum voor Wiskunde en Informatica, Amsterdam,
ftp.cwi.nl/pub/CONTENT.
Kuznetsov, Yu.A., 1998b. Elements of Applied Bifurcation Theory. 2nd
Edition. Springer, New York.
Mahmoud, M.S., Khalil, H.K., 2002. Robustness of high-gain observerbased nonlinear controllers to unmodeled actuators and sensors.
Automatica 38, 361369.
Mangold, M., Kienle, A., Gilles, E.D., Mohl, K.D., 2000.
Nonlinear computation in DIVAmethods and applications. Chemical
Engineering Science 55 (2), 441454.
Marino, R., Tomei, P., 1995. Nonlinear Control Design. Prentice-Hall,
Upper Saddle River, NJ.
Mnnigmann, M., Marquardt, W., 2002a. Bifurcation placement of Hopf
points for stabilization of equilibria. In: Proceedings of the 15th IFAC
World Congress, Barcelona, Spain.
Mnnigmann, M., Marquardt, W., 2000b. Normal vectors on manifolds
of critical points for parametric robustness of equilibrium solutions of
ODE systems. Journal of Nonlinear Science 12, 85112.
Mnnigmann, M., Marquardt, W., 2003. Steady-state process optimization
with guaranteed robust stability and feasibility. AIChE Journal 49 (12),
31103126.
Morari, M., Zariou, E., 1989. Robust Process Control. Prentice-Hall,
Upper Saddle River, NJ.
Paladino, O., Ratto, M., 2000. Robust stability and sensitivity of real
controlled CSTRs. Chemical Engineering Science 55, 321330.
Perko, L., 1996. Differential Equations and Dynamical Systems. 2nd
Edition. Springer, New York.
Recke, B., Andersen, B.R., JZrgensen, S.B., 2000. Bifurcation control of
sample chemical reaction systems. In: Proceedings of the ADCHEM
2000. Pisa, Italy, pp. 575580.
Scott, S.K., 1991. Chemical Chaos International Series of Monographs
on Chemistry. Oxford University Press, New York.
Seydel, R., 1994. Practical Bifurcation and Stability Analysis. Springer,
New York.
Uppal, A., Ray, W.H., Poore, A.B., 1974. On the dynamic behavior of
continuous stirred tank reactors. Chemical Engineering Science 29,
967985.
Vidyasagar, M., 1993. Nonlinear Systems Analysis. second ed. PrenticeHall, NJ.
Zeitz, M., 1987. The extended Luenberger observer for nonlinear systems.
Systems and Control Letters 9, 149156.
Zhang, Y.C., Zamamiri, A.M., Henson, M.A., Hjortso, M.A., 2002. Cell
population models for bifurcation analysis and nonlinear control of
continuous yeast bioreactors. Journal of Process Control 12, 721734.

Das könnte Ihnen auch gefallen