Wang Decoupling Control
Wang Decoupling Control
Introduction
Prior to the 1940s, most industrial systems were run essentially manually
or using on-off control. Many operators were needed to keep watch on the
many variables in the plant. With the continuous industrialization over the
last several decades, manufacturing and production have taken off at an in-
credibly high speed in almost every part of the world. As a consequence of
the expanding scale and volume of production, there could be hundreds or
even thousands of variables to be controlled in a plant. The manual effort
thus needed in operation is tremendous. With increasing labor and equip-
ment costs, and with eager demand for high precision, quality and efficiency,
the idea of employing operators for the control of physical systems rapidly
became uneconomical and infeasible. Automatic control thus becomes a solu-
tion much sought after. The fundamental component in an automatic control
system is the so called “controller”. It could either be a piece of hardware
or software code in a computer. Its job is to receive information about the
system from a variety of sensors, process it, and automatically generate com-
mands for corrective action to bring the variable of interest to its desired
value or trajectory. Wide applications of automatic control have driven great
attention to its theoretical development. Since the age of the familiar clas-
sical control theory, many new and sophisticated theories have evolved and
striking developments have taken place especially since 1950s. Optimal and
stochastic control, state-space theory and adaptive control are only a few to
name. The advances in microelectronics which give cheap microprocessors
whose computational power is continuously increasing have also acted as a
supportive catalyst in the development of new concepts by providing conve-
nient test grounds and platforms for extensive simulations to be performed
in the various stages of design and verification, and enabling effective im-
plementation of advanced control algorithms. Today, automatic controllers
have found a much wider scope of usage than only in industrial processes. It
may range from missile tracking in military applications, down to water and
temperature control in washing machines. Automatic control systems have
permeated life in all advanced societies.
are limited to 30%, then each off-diagonal element will have relative gain less
than 3%, so the system is almost decoupled.
It should be, however, pointed out that there are cases where decoupling
should not be used, instead, couplings are deliberately employed to boost
performance such as airplane control systems. The extreme case is that de-
coupling cannot be used at all since it may destabilize the system. In short,
decoupling is a common practice in industry, but needs to be justified from
the real situation and requirements before use.
The study of decoupling linear time-invariant multivariable systems has
received considerable attentions in both control theory and industrial prac-
tice for several decades, dating back at least to (Voznesenskii, 1938). Dur-
ing that early times, this problem was treated with transfer matrices, see
(Kavanagh, 1966; Tsien, 1954) for reviews. The state space approach to de-
coupling was initiated by Morgan (Morgan, 1964). A seminal paper by Falb
and Wolovich (Falb and Wolovich, 1967) presented a necessary and suffi-
cient condition on the solvability of decoupling problem for square systems
described by state space models. The equivalent condition for systems de-
scribed by transfer matrices was given by Gilbert (Gilbert, 1969) later. In
these investigations, the problem is confined to the case of scalar input and
output blocks and hence with equal number of inputs and outputs. The more
general block decoupling problem was first defined and solved by Wonham
and Morse (Morse and Wonham, 1970; Wonham and Morse, 1970) using ge-
ometric approach, see also the monograph (Wonham, 1986). For this kind
of general block decoupling problem, an alternative approach was developed
by Silverman and Payne (Silverman and Payne, 1971) based on structure
algorithm proposed by Silverman (Silverman, 1969). In these works, it is
still required that the transformation matrix between the new control in-
puts and old control inputs is nonsingular. The complete solution to gen-
eral Morgan problem, without any assumptions on either system matrices
or feedback matrices, is due to Descusse et al. (Descusse, 1991; Descusse et
al., 1988). Other developments can be found in (Commault et al., 1991; Des-
oer and Gundes, 1986; Hautus and Heymann, 1983; Koussiouris, 1979; Kous-
siouris, 1980; Pernebo, 1981; Williams and Antsaklis, 1986), to name a few.
A parallel avenue for the study of decoupling problem is for unity out-
put feedback systems based on input-output models. The existence condition
for decoupling by unity output feedback compensation is that the plant has
full row rank, which is very simple and well known. But the problem of de-
coupling in such a configuration while maintaining internal stability of the
decoupled system appears to be much more difficult. Under the assump-
tion that the plant has no unstable poles coinciding with its zeros, the di-
agonal decoupling problem is solvable and a parameterization of controllers
solving the problem could be obtained (Safonov and Chen, 1982; Vardu-
lakis, 1987; Peng, 1990; Desoer, 1990; Lin and Hsieh, 1991). The assumption
is however not necessary and is relaxed later by Linnemann and Maier (1990)
for 2 × 2 plants, and by Wang (1992) for square plants, along with a con-
4 1. Introduction
It is well known (Astrom and Hagglund, 1995) that the attenuation of load
disturbance is of a primary concern for any control system design, and is even
the ultimate objective for process control, where the set-point may be kept
unchanged for years (Luyben, 1990). In fact, countermeasure of disturbance
is one of the top factors for successful and failed applications (Takatsu and
Itoh, 1999). If the disturbance is measurable, feedforward control is a useful
tool for cancelling its effect on the system output (Ogata, 1997), while feed-
back control may still be used as usual for stability and robustness. Formally,
the problem of (dynamic) disturbance decoupling is to find a compensator
such that the resulting closed-loop transfer function from the disturbance
to the controlled output is equal to zero. This problem is well-known and
6 1. Introduction
has been investigated extensively. It was actually the starting point for the
development of a geometric approach to systems theory. Numerous investiga-
tors, employing a variety of mathematical formulations and techniques, have
studied this problem.
In a geometric setting the problem using a state feedback was consid-
ered by Basile and Marro (1969) and Wonham (1986). Equivalent frequency
domain solvability conditions were given by Bhattacharyya (1980), Bhat-
tacharyya (1982). It is more realistic to assume that only output is available
for feedback. In this context, the problem was solved by Akashi and Imai
(1979) and Schumacher (1980). However, they did not consider the issue
of stability. The disturbance decoupling by output feedback with stability
or pole placement was solved for the first time by Willems and Commault
(1981). This problem was also considered from an input-output viewpoint by
Kucera (1983) for single variable system and by Pernebo (1981) for multi-
variable systems. The structure of the control system which decouples the
disturbances is now well understood in both geometric terms and frequency
domain terms.
To reject unmeasurable disturbances which are more often encountered in
industry, one possible way is to rely on the single controller in the feedback
system in addition to normal performance requirements. Then, there will be
inevitably a design trade-off between the set-point response and disturbance
rejection performance (Astrom and Hagglund, 1995). To alleviate this prob-
lem, a control scheme which introduces some add-on mechanism to the con-
ventional feedback system was introduced, first in the filed of the servo control
in mechatronics, and is called the disturbance observer (Ohnishi, 1987). It
was further refined by Umeno and Hori (1991). The disturbance observer
estimates the equivalent disturbance as the difference between the actual
process output and the output of the nominal model. The estimate is then
fed to a process inverse model to cancel the disturbance effect on the output.
The disturbance observer makes use of the plant inverse, and its applica-
tion needs some approximation as the plant inverse is usually not physically
un-realizable due to improperness and/or time delay.
The complete disturbance decoupling is usually difficult to achieve. A
more realistic requirement will be static or asymptotic disturbance decou-
pling. The latter problem falls in a general framework of asymptotic tracking
and regulation problem, a fundamental control system problem. Wonham
(1986) gives a complete exposition, where solutions together with (necessary
and sufficient) solvability conditions, and a procedure for the construction
of a compensator are obtained. This problem had been alternatively been
considered by Davison (1975) in an algebraic setting. Both works are carried
out in a state-space viewpoint. This problem, which can also be formulated
in a frequency domain setting, was also approached from an input-output
viewpoint such as Francis (1977) who considered the case where feedback
signals coincide with the outputs; Chen and Pearson (1978) who considered
the case where feedback signals are certain functions of outputs; and finally
1.2 Disturbance Decoupling 7
Pernebo (1981) who considered the general case where feedback signals may
be different from outputs and not related to each other. The crucial condition
for the solvability in frequency domain setting is the existence condition for
the solution of Diophantine equation (Kucera, 1979), which, in a special case,
is simplified to skew-primeness (Wolovich, 1978). Now, the problem has been
well solved in both state space and frequency domains. The relevant solutions
in terms of the internal model principle are popular in control society.
In practice, we often encounter the situation where the reference com-
mands to be tracked and/or disturbance to be rejected are periodic signals,
e.g., repetitive commands or operations for mechanical systems such as in-
dustrial robots, and disturbances depending on the frequency of the power
supply (Hara et al., 1988). Disturbances acting on the track-following servo
system of an optical disk drive inherently contain significant periodic compo-
nents that cause tracking errors of a periodic nature. For such disturbances,
the controller designed for step type reference tracking and/or disturbance
rejection will inevitably give an uncompensated error. Though the internal
model principle seems applicable, the unstable modes of a periodic distur-
bance is infinite and no longer rational, and actually it contains a time delay
term in its denominator. For such delay and unstable systems, no stabilizing
scheme could be found in the literature. Therefore, one has to abandon the in-
ternal model principle and seek other solutions. Hara et al. (1988) pioneered a
method for periodic disturbance rejection, which is now called repetitive con-
trol. However, this method potentially makes the closed-loop system prone
to instability, because the internal positive feedback loop that generates a
periodic signal reduces the stability margin. Consequently, the trade-off be-
tween system stability and disturbance rejection is an important yet difficult
aspect of the repetitive control system design. Moon et al. (1998) proposed
a repetitive controller design method for periodic disturbance rejection with
uncertain plant coefficients. The design is performed by analyzing the fre-
quency domain properties, and Nyquist plots play a central role throughout
the design phase. It is noted that such a method is robust at the expense
of performance deterioration compared with the nominal case, and further,
the disturbance cannot be fully compensated for even under the ideal case.
Another way to solve the periodic disturbance problem is to use the double
controller scheme (Tian and Gao, 1998). However, the complexity and the
lack of tuning rules hinder its application. A plug-in adaptive controller (Hu
and Tomizuka, 1993; Miyamoto et al., 1999) is proposed to reject periodic
disturbances. An appealing feature is that turning on or off the plug module
will not affect the original structure. However, the shortcoming of this method
is that the analysis and implementations are somewhat more complex than
the conventional model based algorithm. An alternative scheme, called the
virtual feedforward control, is proposed in Zhang (2000) for asymptotic re-
jection of periodic disturbance. The periodic disturbance is estimated when
a steady state periodic error is detected, and the virtual feedforward control
is then activated to compensate for such a disturbance.
8 1. Introduction
The book assumes the pre-requisite of basic linear system theory from readers.
It is organized as follows.
Chapter 2 reviews some notions and results on linear systems. Multivari-
able linear dynamic systems and their representations are introduced. Poly-
nomial and rational function matrices are studied in details. Matrix fraction
descriptions and multivariable pole and zeros are covered. General methods
for model reduction are highlighted and two special algorithms taking care
of stability of reduced models are presented. Popular formulas for conver-
sions between continuous time and discrete time systems are discussed and
a technique with high accuracy and stability is described.
Chapter 3 considers stability and robustness of linear feedback systems.
Internal stability of general interconnected systems is addressed and a pow-
erful stability condition is derived which is applicable to systems with any
feedback and/or feedforward combinations. A special attention is paid to the
conventional unity output feedback configuration and the simplifying condi-
tions and Nyquist-like criteria are given. The plant uncertainties and feedback
system robustness are introduced.
Chapter 4 considers decoupling problem by state feedback. Necessary and
sufficient solvability conditions are derived and formulas for calculating feed-
back and feedforward gain matrices given. Pole and zeros of decoupled sys-
tems are discussed. Geometric methods are not covered.
Chapter 5 addresses decoupling problem by unity output feedback com-
pensation. The polynomial matrix approach is adopted. The diagonal de-
coupling problem for square plants is first solved and then extended to the
general block decoupling case for non-square plants. A unified and indepen-
dent solution is also presented. In all the cases, stability is included in the
discussion, the necessary and sufficient condition for solvability is given, and
the set of all the compensators solving the problem is characterized.
Chapter 6 discusses decoupling problem for plants with time delay. Con-
ventional unity output feedback configuration is used exclusively. Our empha-
sis is to develop decoupling methodologies with necessary theoretical supports
as well as the controller design details for possible practical applications. The
new decoupling equations are derived in a transfer function matrix setting,
and achievable performance after decoupling is analyzed where characteriza-
tion of the unavoidable time delays and non-minimum phase zeros that are
inherent in a feedback loop is given. An effective decoupling control design
method is then presented with stability and robustness analysis.
Chapter 7 considers decoupling problem in connection with time delay
compensation. The presence of time delay in a feedback control loop could
impose serious control performance limitations, and they can be released
only by time delay compensation. The Smith control scheme and the internal
model control (IMC) structure are popular for SISO time delay compensation,
and are extended to MIMO delay systems in this chapter.
1.3 Organization of the Book 9
In this chapter, we will review some preliminary notions and results on linear
systems that will be needed subsequently for the solution of various decou-
pling problems in later chapters. Multivariable linear dynamic systems and
their representations are introduced. Polynomial and rational function ma-
trices are studied in details. Matrix fraction descriptions and multivariable
poles and zeros are covered. State-space realizations from matrix fraction
descriptions are discussed. Model reduction is important in many ways, ex-
isting methods are highlighted and two algorithms taking care of stability
of reduced models are presented. Conversions between continuous time and
discrete time systems are reviewed and a technique with high accuracy and
stability preservation is described.
where L and L−1 stand for the Laplace and inverse Laplace transforms,
respectively.
In the case of u(t) = 0, ∀t ≥ t0 , it is easy to see from the solution (2.3)
that for any t1 ≥ t0 and t ≥ t0 , we have
Theorem 2.1.1. The system (2.1) is controllable if and only if one of the
following conditions is true:
i) The controllability matrix:
Qc := B AB A2 B · · · An−1 B ,
has full row rank (over the real number field R);
ii) For α ∈ Rn , αT eAt B = 0 implies α = 0 for any t > 0, i.e, the n rows of
eAt B are linearly independent (over the real number field R);
iii) The matrix [sI − A, B] has full row rank for any s ∈ C, the complex
number field;
iv) The eigenvalues of A − BK can be freely assigned (with the restriction
that complex eigenvalues are in conjugate pairs) by a suitable choice of
K.
Note that the unforced solution x(t) to the stable system (2.1), i.e., under
u(t) ≡ 0, meets x(t) = eAt x(0) → 0 when t → ∞.
Definition 2.1.3. The dynamical system (2.1), or the pair (A, B), is said
to be state-feedback stabilizable if there exists a state feedback u = −Kx such
that the feedback system,
ẋ = (A − Bk)x,
is stable.
Taking the Laplace transform of the system in (2.1) and (2.2) under the
zero initial condition x(0) = 0 yields the transfer function matrix of the
system:
On the other hand, given the transfer function matrix G(s), the system in
(2.1) and (2.2) is said to be a (state-space) realization of G(s) if (2.7) holds.
The realization is called minimal if (A, B) is controllable and (A, C) is ob-
servable.
The computation of the transfer matrix G(s) from the state-space model
can be carried out with the Faddeev algorithm as follows. Let the character-
istic polynomial of the n × n matrix A be denoted by
Then, write
1
(sI − A)−1 = (Γn−1 sn−1 + Γn−2 sn−2 + · · · + Γ0 ), (2.9)
φ(s)
where αn−1 , αn−2 , . . . , α0 , Γn−1 , Γn−2 , · · · , Γ0 can be calculated in a recursive
manner from
Γn−1 = I, αn−1 = −tr (AΓn−1 ),
Γn−2 = AΓn−1 + αn−1 I, αn−2 = −tr (AΓn−2 )/2,
Γn−3 = AΓn−2 + αn−2 I, αn−3 = −tr (AΓn−3 )/3,
··· ···
Γi = AΓi+1 + αi+1 I, αi = −tr (AΓi )/(n − i),
··· ···
Γ0 = AΓ1 + α1 I, α0 = −tr (AΓ0 )/n,
0 = AΓ0 + α0 I,
where tr (X), the trace of X, is the sum of all the diagonal elements of the
matrix X.
14 2. Representations of Linear Dynamic Systems
Γ2 s2 + Γ1 s + Γ0
(sI − A)−1 = ,
φ(s)
where
100 9 −2 0 20 −10 0
Γ2 = 0 1 0 , Γ1 = −3 6 0 , Γ0 = −15 5 0 ,
001 0 0 5 0 0 −2
and
φ(s) = s3 + 10s2 + 23s − 10,
which in turn lead to
1
C(sI − A)−1 B = {CΓ2 Bs2 + CΓ1 Bs + CΓ0 B}
φ(s)
1 01 2 −2 9 −10 20
= s + s+
φ(s) 00 0 0 0 0
2
1 −2s − 10 s + 9s + 20
= ,
φ(s) 0 0
and
G(s) = C(sI − A)−1 B + D
" #
−2s−10 s2 +9s+20
+ 1
= φ(s) φ(s) . ♦
0 1
The transfer function matrices for MIMO systems without time delay
are in fact rational function matrices of complex variable, namely, matrices
whose generic element is a rational function, a ratio of polynomials with real
coefficients. A transfer function matrix is said to be proper when each element
is a proper rational function, i.e.,
where the notation K < ∞ means that each element of matrix K is finite.
Analogously, G(s) is strictly proper if
2.2 Polynomial Matrices 15
lim G(s) = 0.
s→∞
is such an example. Taking the Laplace transform under zero initial condition
yields
R(s) is called the field of (real) rational functions. A matrix P (s) whose
elements are polynomials is called a polynomial matrix. The set of p × m
16 2. Representations of Linear Dynamic Systems
polynomial matrices is denoted by Rp×m [s]. Similarly, Let Rp×m (s) denote
the set of p × m matrices with elements in R(s). A matrix T (s) ∈ Rp×m (s)
is called a (real) rational function matrix. The rank of a polynomial or ratio-
nal function matrix is the maximum number of its linearly independent row
vectors or its column vectors over the field R(s) and it is denoted by rank
T (s).
Let P (s) ∈ Rm×m [s] be a square polynomial matrix. In general, its determi-
nant, denoted by det P (s), is a polynomial. The following two special cases
deserve attention.
Note that if P(s) is nonsingular, then its inverse P −1 (s) exists and is in
general a rational function matrix. In the case of unimodular P (s), however,
the inversion formula, P −1 (s) = adj(P (s))/ det P (s), tells us that P −1 (s) be-
comes a polynomial matrix, and is also unimodular due to
det P (s) det P −1 (s) = 1. In fact, P (s) is unimodular if and only if P (s) and
P −1 (s) are both polynomial matrices.
Unimodular matrices are closely related to elementary operations. To be
precise, elementary row and column operations on any polynomial matrix
P (s) ∈ Rp×m [s] are defined as follows:
i) interchange any two rows or columns of P(s);
ii) multiply row or column i of P(s) by a non-zero real number;
iii) add to row or column i of P(s) a polynomial multiple of row or column
j, j 6= i.
These elementary operations can be accomplished by multiplying the given
P (s) on the left or on the right by elementary unimodular matrices, namely
matrices obtained by performing the above elementary operations on the
identity matrix I. It can also be shown that every unimodular matrix may
be represented as the product of a finite number of elementary matrices.
(i) for a(s), b(s) ∈ R[s] such that a(s)b(s) 6= 0, there holds ∂(a(s)b(s)) ≥
∂(a(s));
2.2 Polynomial Matrices 17
(ii) the polynomial division theorem: for every a(s), b(s) ∈ R[s], b(s) 6= 0,
there exists two elements q(s), r(s) ∈ R[s] such that a(s) = q(s)b(s)+r(s)
and either r(s) = 0 or ∂r(s) < ∂b(s).
We now extend these facts to the matrix case.
Consider a p × m polynomial matrix P(s). The i − th column (resp.
row) degree of P(s) is defined as the highest degree of all m polynomials
in the i − th column (resp. row) of P (s), and denoted by ∂ci (P (s)) (resp.
∂ri (P (s))). Let ∂ci P (s) = µi , i = 1, 2, . . . , m; ∂ri P (s) = νi , i = 1, 2, . . . , p;
Sc (s) = diag{sµ1 , sµ2 , · · · , sµm } and Sr (s) = diag{sν1 , sν2 , · · · , sνp }. P (s)
can be expressed by
P (s) = Phc Sc (s) + Lc (s)
= Sr (s)Phr + Lr (s),
where Phc (resp. Phr ) is the highest column (resp. row) degree coefficient
matrix of P (s), and Lc (s) (resp. Lr (s)) has strictly lower column (resp. row)
degrees than ∂ci P (s) (resp. ∂ri P (s)).
In the case of square matrices, p = m, it readily follows from linear algebra
that
det P (s) = det Phc · sµ + lower − degree terms in s
and
det P (s) = sν det Phr + lower − degree terms in s
m
X m
X
with µ = µi and ν = νi , which imply, respectively
i=1 i=1
m
X
∂ detP (s) ≤ µi ,
i=1
and
m
X
∂ detP (s) ≤ νi .
i=1
One writes
18 2. Representations of Linear Dynamic Systems
11 s2 0 1 0
+
P (s) = 00 0 s2 2s s
Phc Sc (s) Lc (s)
s2 0 11 10
+ .
= 0 s 21 00
Sr (s) Phr Lr (s)
is column-reduced, and
is row-reduced.
Continue Example 2.2.1, where P (s) is not column-reduced. To reduce
column degrees, we examine just the highest-degree terms in each column:
2 2
s s
,
0 0
which can be transformed to
2
0s
0 0
by a post-multiplication with
1 0
UR (s) = .
−1 1
Thus, we perform
2
s2 + 1 s2 1 0 1s
P (s)UR (s) = =
2s s −1 1 s s
01 s 0 10
+ ,
= 10 0 s2 0s
P̄hc S̄c (s) L̄c (s)
which is already column-reduced.
2.2 Polynomial Matrices 19
Consider now two polynomial matrices N (s) ∈ Rp×m [s] and D(s) ∈
m×m
R [s] with D(s) nonsingular. In general, N (s)D−1 (s) will be a ratio-
nal function matrix. By carrying out polynomial divisions in each element of
N (s)D−1 (s), we have
where Q(s) ∈ Rp×m [s] is polynomial and W (s) ∈ Rp×m (s) is strictly proper.
It can be re-written as
where the left-hand side is polynomial and the right-hand side is strictly
proper. For (2.11) to hold true, both sides must be zero. Hence Q(s) = Q̄(s)
and R(s) = R̄(s).
Proposition 2.2.2. Let D(s) ∈ Rm×m [s] be nonsingular. Then, for any
N (s) ∈ Rp×m [s] there exist unique Q(s), R(s) ∈ Rp×m [s] such that (2.10)
holds true with R(s)D(s)−1 ∈ Rp×m (s) strictly proper.
Analogously, given a nonsingular A(s) ∈ Rp×p [s], then for any B(s) ∈
p×m
R [s] there exist unique Q(s), R(s) ∈ Rp×m [s] such that
Given that
−1 1 1 −1
D (s) = ,
s+1 1 s
we obtain
20 2. Representations of Linear Dynamic Systems
s+1 s 1 0 0 0
1 s2 s3 = s − 1 s2 − s + 1 + 1 1 −1 .
N (s)D−1 (s) =
s+1 s+1
−1 1 0 0 −1 1
Hence,
1 0 0 0 0 0
1 s 1
Q(s) = s − 1 s2 − s + 1 , R(s) = 1 −1 = 1 0 . ♦
s+1 −1 1
0 0 −1 1 −1 0
" µ−1
#
X
Qi−µ = Ni − Qi−j Dj − Ri Dµ−1 , i = γ + µ, γ + µ − 1, · · · , 2, 1, 0,
j=0
where
Ni = 0, i > l,
Qi = 0, i > γ or i < 0,
Ri = 0, i > µ − 1.
Then we have
2.2 Polynomial Matrices 21
Qi = 0, i > l − µ; (2.17)
µ−1
X
Qi = Ni+µ − Qi+µ−j Dj Dµ−1 , i = l − µ, l − µ − 1, . . . , 1, 0; (2.18)
j=0
i
X
Ri = N i − Qi−j Dj , i = µ − 1, µ − 2, · · · , 1, 0. (2.19)
j=0
Q(s) = Q1 s + Q0 .
−1 −4
Q0 = N2 − Q1 D1 D2−1 = 0 −4 .
1 1
22 2. Representations of Linear Dynamic Systems
0 4
R0 = N0 − Q0 D0 = 0 −8 .
0 −1
Finally, we obtain
s − 1 −4
Q(s) = Q1 s + Q0 = 0 −4 ,
1 1
12 −3s + 4
R(s) = (R1 s−1 + R0 s−2 )Sc (s) = 10 −11s + 8 . ♦
−2 −s − 1
Then, we have
P̄ := Dr (s)P (s) = Im sν + Pν−1 sν−1 + Pν−2 sν−2 + · · · + P1 s + P0 , (2.21)
det(P̄ (s)) = det(Dr (s)) · det(P (s)) = s∆ν · det(P (s)), (2.22)
2.2 Polynomial Matrices 23
m
X
where ∆ν = (ν − νi ). Introduce the constant matrix A as follows:
i=1
0 0 −P0
Im −P1
A= . .. . (2.23)
.. .
0 Im −Pν−1
If P (0) is singular, det P (0) is zero, and P (s) has a root at the origin so
that P (s) is unstable. On the other hand, if P (0) is nonsingular, det P (0)
is non-zero so that det P (s) has no roots at the origin. Then, by (2.22), the
non-zero roots of det P (s) will be the same as those of det P (s). We have
actually established the following theorem.
Then, we form
0 0 −4 −2
0 0 0 0
A=
1 0 −4 −1 ,
0 1 −1 −2
and obtain its non-zero eigenvalues as −1, −2, −3, which are all in C− . By
Theorem 2.2.1, P (s) is thus stable. ♦
24 2. Representations of Linear Dynamic Systems
Two polynomial matrices P1 (s), P2 (s) ∈ Rp×m [s] are called equivalent (in C)
if there exist unimodular matrices UL (s) ∈ Rp×p [s], UR (s) ∈ Rm×m [s] such
that
Subtracting the i − th row by qi1 times 1st row, we actually replace p̄i1 (s) by
ri1 (s). Among nonzero ri1 (s), choose one with the lowest degree and bring
it to position (1,1) by suitable interchange of the rows. Continuing this pro-
cedure until all elements in the first column are zero except one at position
(1,1). Now consider the second column of the resulting matrix and, temporar-
ily ignoring the first row, repeat the above procedure until all the elements
below the (2,2) element are zero. If the (1,2) element does not have lower
degree than the (2,2) element, the division algorithm and an elementary row
operation can be used to replace the (1,2) element by a polynomial of lower
degree than the diagonal or (2,2) element. Continuing this procedure with
the third column, fourth column, and so on finally gives the desired Hermite
row form.
2.2 Polynomial Matrices 25
Remark 2.2.1. By interchanging the roles of rows and columns, one can ob-
tain a similar Hermite column form, the details of which we shall not spell
out.
Example 2.2.5. We proceed as follows:
s 3s + 1 1 −(s2 + 2s − 1)
−1 s2 + s − 2 −r−→
3 ↔r1
−1 s2 + s − 2
2
−1 s + 2s − 1 s 3s + 1
r1 →r2 1 −(s2 + 2s − 1) 1 −(s2 + 2s − 1)
−−
−sr− −→
1 →r3
0 −(s + 1)
−r2 →r2
−→ 0 (s + 1)
3 2 2
0 s + 2s + 2s + 1 0 (s + 1)(s + s + 1)
(s+1)r2 →r1 1 2
−−−2−−−−−−−→
−(s +s+1)r2 →r3 0 s + 1 := H(s)
0 0
0 −(s + 1) s
= 0 −1 1 := U (s). ♦
2 2
1 (s + s + 1) −(s + 1)
λi (s)|λi+1 (s), i = 1, 2, · · · , r − 1,
i.e., λi (s) divides λi+1 (s) without remainder. λi (s) are called invariant poly-
nomials of P(s). Moreover, define the determinant divisors of P (s) by ∆i (s)=
26 2. Representations of Linear Dynamic Systems
Proof. Bring the least degree element of P (s) to the (1,1) position. As in
construction of the Hermite forms, by elementary row and column operations,
make all other elements in the first row and column zero. If this (1,1) element
does not divide every other element in the matrix, using the polynomial
division algorithm and row and column interchanges, we can bring a lower-
degree element to the (1,1) position, and then repeat the above steps to zero
all other elements in the first row and column. Then we reach
λ1 (s) 0 · · · 0
0
U 1 P V1 = . ,
.. P1 (s)
0
where λ1 divides every element of P1 . Now repeat the above procedure on
P1 . Proceeding in this way gives the Smith form Λ(s).
Example 2.2.6. Continue Example 2.2.5. With the additional elementary col-
umn operation corresponding to
1 −2
:= UR (s),
0 1
we get the Smith form:
1 0
H(s)UR (s) = 0 s + 1 := Λ(s). ♦
0 0
Example 2.2.7. Consider a polynomial matrix,
1 −1
P (s) = s2 + s − 4 2s2 − s − 8 .
s2 − 4 2s2 − 8
One proceeds,
∆0 (s) = 1,
∆1 (s) = GCD{1, −1, s2 − s − 4, 2s2 − s − 8, s2 − 4, 2s2 − 8} = 1,
ü ü ü ü
ü 1 −1 ü ü 1 −1 üü
∆2 (s) = GCD{üü 2 ü , ü ,
s + s − 4 2s2 − s − 8 ü ü s2 − 4 2s2 − 8 ü
ü 2 ü
ü s + s − 4 2s2 − s − 8 ü
ü 2 ü}
ü s −4 2s2 − 8 ü
= (s + 2)(s − 2).
2.2 Polynomial Matrices 27
It follows that
∆1 ∆2
λ1 (s) = = 1, λ2 (s) = = (s + 2)(s − 2).
∆0 ∆1
Thus, one obtains the Smith form:
1 0
Λ(s) = 0 (s + 2)(s − 2) . ♦
0 0
Definition 2.2.3. Let P (s), L(s), and R(s) be polynomial matrices such that
P (s) = L(s)R(s).
Then R(s) (resp. L(s)) is called a right (resp. left) divisor of P (s), and P (s)
a left (resp. right) multiple of R(s) (resp. L(s)).
1 0 1 0
R3 (s) = , R4 = ,
0 s+1 0 s(s + 1)
Definition 2.2.4. Let two polynomial matrices N (s) and D(s) have the
same number of columns.
28 2. Representations of Linear Dynamic Systems
N (s) = Ñ (s)R(s),
D(s) = D̃(s)R(s);
(ii) R(s) is called a greatest common right divisor (GCRD) of N (s) and D(s)
if R(s) is a left multiple of any other common right divisor R̄(s) of N (s)
and D(s), i.e. there is a polynomial matrix L(s) such that
R(s) = L(s)R̄(s);
(iii) N (s) and D(s) are called right coprime if all common right divisions of
N (s) and D(s) are unimodular.
Analogously, we can define greatest common left divisors (GCLD) and left
coprimeness of two polynomial matrices having the same number of rows.
Greatest common divisors can be found by elementary operations on a com-
posite matrix formed by N (s) and D(s), as indicated by the following theo-
rem.
Theorem 2.2.4. Let N (s) ∈ Rp×m [s] and D(s) ∈ Rm×m [s].
(i) If a unimodular matrix U (s) ∈ R(p+m)×(p×m) [s] and a square polyno-
mial matrices R(s) ∈ Rm×m[s] are such that
D(s) R(s)
U (s) = . (2.26)
N (s) 0
then all GCRDs of N (s) and D(s) must be nonsingular and can differ
only by unimodular (left) divisors. In particular, if a GCRD is unimod-
ular, then all GCRDs must be unimodular.
where U11 (s) ∈ Rm×m [s] and U22 (s) ∈ Rp×p [s]. Then the polynomial matrix
U −1 (s) can be partitioned similarly as
2.2 Polynomial Matrices 29
−1
U11 (s) U12 (s) V11 (s) V12 (s)
= .
U21 (s) U22 (s) V21 (s) V22 (s)
It follows that
D(s) V11 (s) V12 (s) R(s)
= ,
N (s) V21 (s) V22 (s) 0
or equivalently
indicating that R(s) is a common right divisor of D(s) and N (s). Notice next
from (2.26) that
Notice that the reduction of [DT (s) N T (s)]T to [RT (s) 0]T can be
achieved by elementary row operations such that the last p rows from the
bottom on the right-hand side become zero. In particular, we could use the
procedure for getting the Hermite row form, but this is more than necessary.
It follows from Example 2.2.5 that there is a unimodular matrix U (s) such
that
30 2. Representations of Linear Dynamic Systems
1 2
D(s) 0 s + 1
U (s) · · · =
··· ··· .
N (s)
0 0
where X(s) := R−1 (s)X̂(s) and Y (s) := R−1 (s)Ŷ (s) are both polynomial.
Conversely, suppose that there exist two matrices X(s) and Y (s) satisfying
(2.29). Let R(s) be a GCRD of N (s) and D(s), i.e.
yielding
is of rank defect, and then D(s) and N (s) are not coprime. One may write
2
s + 2s s + 3 s+2 s+3
D(s) 2 s0
= 2s − s 3s − 2 = 2s − 1 3s − 2
N (s) 01
s 1 1 1
D̃(s)
:= R(s).
Ñ (s)
and find that the right divisor R(s) is not unimodular as detR(s) = s =
6 const,
which is in agreement with non-coprimeness of D(s) and N (s). One notes that
h iT
the first and third rows of D̃T (s) Ñ T (s) form a nonsingular matrix for
any s ∈ C. Thus, D̃(s) and Ñ (s) are right coprime and R(s) is actually a
GCRD of them. ♦
Nd (s) = d(s)G(s)
The degrees of these two PMFs are different in general, and it should not be
surprising that lower-degree PMFs can be found if some effort is invested.
Let dci (s) (resp. dri (s)) be a least common denominator of all elements in
the ith column (resp. ith row) of G(s), then
and
In the single-input and single output case, the issue of common factors in the
numerator and denominator polynomials of G(s) draws attention. The utility
of PMFs begins to emerge from the corresponding concept in the matrix case.
For PMFs, one of the polynomial matrices is always nonsingular, so only
nonsingular common divisors occur. Suppose that G(s) is given by a right
PMF:
and that R(s) is a common right divisor of N (s) and D(s). Then
e (s) = N (s)R−1 (s)
N and e
D(s) = D(s)R−1 (s) (2.31)
are polynomial matrices, and they provide another right PMF for G(s) since
e (s)D
N e −1 (s) = N (s)R−1 (s)R(s)D−1 (s) = G(s).
The degree of this new PMF is no greater than the degree of the original,
since
e
deg [det D(s)] = deg [det D(s)] + deg [det R(s)].
Proof. Given a G(s), it is always possible, say by (2.30), to get a right PMF,
G(s) = N (s)D−1 (s). Using Theorem 2.2.4, we can extract a GCRD, R(s), of
N (s) and D(s) from
D(s) R(s)
U (s) = . (2.32)
N (s) 0
−1
A right coprime PMF is then obtained as NR (s)DR (s) where
Next, for any two right coprime PMFs, G(s) = N (s)D−1 (s) = N1 (s)D1−1 (s),
by Theorem 2.2.5, there exist polynomial matrices X(s), Y (s), A(s), and
B(s) such that
34 2. Representations of Linear Dynamic Systems
and
Since N (s)D−1 (s) = N1 (s)D1−1 (s), we have N1 (s) = N (s)D−1 (s)D1 (s). Sub-
stituting this into (2.33) gives
or
One sees that both D1−1 D(s) and D−1 (s)D1 (s) are polynomial matrices, but
they are inverses of each other. Thus, both must be unimodular. Let
Then, we have
and U −1 (s) as
DR −XL
U −1 (s) = .
NR YL
Theorem 2.3.2. Let N (s)D−1 (s) is a right PMF of G(s) with the operations
−1
leading to (2.32) performed. Then NR DR is a right coprime PMF of G(s),
−1
and DL NL is a left coprime PMF of G(s).
2.3 Rational Matrices and Polynomial Matrix Fractions 35
YR DR + XR NR = I,
−1
where YR and XR are both polynomial. Thus NR DR is a right coprime MFD
of G(s).
The (2,1) block of (2.35) gives
D L NR = NL D R . (2.36)
We now show nonsingularity of DL by contradiction. Assume that there is
a non-zero polynomial vector α(s) such that αT DL = 0. Then by (2.36),
one has 0 = αT DL NR = αT NL DR . This implies αT NL = 0 since DR is
nonsingular. It then follows that
αT [NL XL + DL YL ] = αT NL XL + αT DL YL = 0.
But the (2,2) block of (2.35):
NL XL + DL YL = I, (2.37)
shows
αT [NL XL + DL YL ] = αT I = αT 6= 0.
−1 −1 −1
From (2.36), DL NL = NR D R = G(s), so that DL NL is a left PMF of
G(s), and it is coprime in view of (2.37).
It follows from Example 2.2.5 that U which makes U [DT N T ]T = [RT 0]T
is
0 −(s + 1) s
U = 0 −1 1 ,
2 2
1 (s + s + 1) −(s + 1)
−1
from which we get a left coprime PMF, DL (s)NL (s), of G(s) where
36 2. Representations of Linear Dynamic Systems
Proof. Since NR and DR are right coprime, there are polynomials XR and
YR such that YR DR + XR NR = I. By Theorem 2.3.2, we can have a left
−1 −1
coprime polynomial matrices NL and DL such that DL NL = NR D R . The
left coprimeness of NL and DL implies the existence of polynomials X̃L and
eL = I. Writing these relations in matrix form gives
ỸL such that DL YeL +NL X
" #
YR XR DR −X eL IQ
= ,
−NL DL NR YeL 0 I
G = UL M U R ,
and
1 /ψ1 0
2 /ψ2
..
M = . ∈ Rp×m (s) (2.39)
r /ψr
0 0
is the so called Smith-McMillan form of G(s), where (i , ψi ) are pairs of
monic polynomials copime to each other, satisfying the following divisibility
properties
ψi+1 |ψi , i = 1, 2, · · · , r − 1,
i |i+1 , i = 1, 2, · · · , r − 1,
It is written as
1
G(s) = P (s),
(s + 1)(s + 2)
Definition 2.3.2. (poles and zeroes) Let G(s) be a rational matrix with
the Smith-McMillan form M (s) in (2.39), define the pole (or characteristic)
polynomial and zero polynomial of G(s), respectively, as
p(s) = ψ1 (s)ψ2 (s) · · · ψr (s),
z(s) = ε1 (s)ε2 (s) · · · εr (s).
The degree of p(s) is called the McMillan degree of G(s).
Example 2.3.4. The pole and zero polynomials of G(s) in Example 2.3.3 are
p(s) = (s + 1)2 (s + 2),
z(s) = (s − 2).
Hence G(s) has poles {−1, −1, −2} and a zero {2}. ♦
Corollary 2.3.1. Let G(s) ∈ Rp×m (s) have rank r, and N (s) and D(s) be
the numerator and denominator of any coprime (right or left) PMF of G(s),
respectively. Then,
(i) ρ is a pole of G(s) if and only if detD(ρ) = 0;
(ii) ζ is a zero of G(s) if and only if rank N (ζ) < r.
In view of the above corollary, one can find poles and zeros of G(s) from
any coprime PMF of it (one may not necessarily get its Smith-McMillan
form). It is easily shown that ρ is a pole of G(s) if and only if it is a pole
of some element of G(s). But one can hardly tell its multiplicity in p(s)
from individual elements. The case is worse for zeros: a zero of elements in
G(s) has no relationship with zeros of G(s). Another interesting property of
multivariable poles and zeros is that a multivariable system can have a pole
and a zero at the same place, but they do not cancel each other. For instance,
1
s+1 0
G= ,
s+1
0 s
Definition 2.3.3. For a rational matrix G(s), if there are polynomial ma-
trices NL , D and NR such that NR D−1 NL = G, then NL D−1 NR is called a
right-left PMF of G. Furthermore, if NR and D are right coprime and D and
NL are left coprime, NR D−1 NL is called a right-left coprime PMF. Obvi-
ously, for NL = I (resp. NR = I), NR D−1 NL reduces to normal right (resp.
left) PMF NR D−1 (resp.NL D−1 ).
Theorem 2.3.6. For a right-left coprime PMF, G = NR D−1 NL , det D is
the pole polynomial of G(s) modulo a nonzero constant.
−1
Proof. Let N L D be a right coprime PMF of D−1 NL , so that D−1 NL =
−1
N L D and detD = detD modulo a nonzero real number. Noting that
(NR N L )(D)−1 = G and Theorem 2.3.5, the proof would be completed if we
show that (NR N L , D) is a right coprime.
Since both pairs, (NR , D) and (N L , D), are coprime, respectively, there
are polynomials X, Y , X and Y such that
XD + Y NR = I, (2.42)
X D + Y N L = I. (2.43)
By postmultiplying (2.42) by D−1 NL , we get successively
XNL + Y NR D−1 NL = D−1 NL ,
−1 −1
XNL + Y NR N L D = N LD ,
XNL D + Y NR N L = N L .
Substituting this into (2.43) yields
(X + Y XNL )D + (Y Y )NR N L = I,
indicating coprimeness of NR N L and D.
Example 2.3.5. Consider a general state-space model:
ẋ(t) = Ax(t) + Bu(t),
y(t) = Cx(t).
Its transfer function matrix is given by
G(s) = C(sI − A)−1 B.
So a state-space realization of the system naturally gives rise to a right-left
PMF of the transfer matrix. Furthermore, if the model is both observable
and controllable or minimal, then by Theorem Theorem 2.1a, (C, sI − A) is a
right coprime, and (sI − A, B) is left coprime, and C(sI − A)−1 B is actually
a right-left coprime fraction of G(s). It follows form Theorem Theorem 2.13a
that det(sI − A) coincides with the pole polynomial of G(s). ♦
2.3 Rational Matrices and Polynomial Matrix Fractions 41
Recall from Section 2.1 that a rational transfer function matrix G(s) is said
to be proper if lims→∞ G(s) < ∞ and strictly proper if lims→∞ G(s) = 0. In
the scalar case, a transfer function is proper if the degree of the numerator
polynomial is less than or equal to the degree of the denominator polynomial.
The situation is not so simple in the matrix case, as we shall now explore.
Lemma 2.3.1. Let G(s) ∈ Rp×m (s) be proper (resp. strictly proper) and
have a right PMF, N (s)D−1 (s). Then, every column of N (s) has degree less
than or equal to (resp. strictly less than) that of the corresponding column of
D(s).
Let now µj = ∂cj D(s), the j-th column degree of D(s), and evaluate
m
X
lim nij (s)s−µj = ( lim gik (s))( lim dkj (s)s−µj ),
s→∞ s→∞ s→∞
k=1
which is finite (resp. zero) if gik is proper (resp. strictly proper). Hence we
have ∂cj N (s) ≤ ∂cj D(s) (resp. ∂cj N (s) < ∂cj D(s)).
The degrees of the columns of N (s) are less than those of the corresponding
columns of D(s). But
h 2 i
N (s)D−1 (s) = s −s+1
s (1 − s)
is not proper.
To obtain necessary and sufficient conditions for the properness of N (s)×
D−1 (s), suppose that D(s) is column-reduced with its highest column degree
coefficient matrix Dhc being nonsingular. Let Sc (s) = diag[sµ1 , sµ2 , ..., sµm ] .
Then one can write
If G(s) ∈ Rp×m (s) is proper, we can have a state space realization (A, B,
C, D) in (2.1-2.2). Obviously, one sees D = G(∞) and G(s) − G(∞) = G(s)
will be strictly proper. We thus aims to get A, B and C from a right PMF
N̄ (s)D̄−1 (s) of a strictly proper G(s). It follows from Proposition 2.2.1 that
there is a unimodular matrix U (s) such that D(s) = D̄(s)U (s) is column-
reduced. So, without loss of generality, we assume that N (s)D−1 (s) is a
right PMF of a p × m strictly proper G(s) with D(s) column-reduced and
∂cj N (s) < ∂cj D(s) = µj , j = 1, 2, · · · , m. Let n = Σµj and write D(s) as
for suitable constant matrices Dhc and Dlc , where Sc (s) = diag{sµ1 , sµ2 , · · · ,
sµm } as before,
Similarly, we write
N (s) = Nc Φc (s)
−1
A = Ao − Bo Dhc Dlc , (2.44)
−1
B= Bo Dhc , (2.45)
C = Nc . (2.46)
We write D(s) as
−1
D(s) = Dhc (Sc (s) + Dhc Dlc Φc (s))
:= Dhc Γ (s),
and look at
−1
(sI − A)Φc (s) = (sI − Ao + Bo Dhc Dlc )Φc (s)
−1
= (sI − Ao )Φc (s) + Bo Dhc Dlc Φc (s)
−1
= Bo Sc (s) + Bo Dhc Dlc Φc (s)
= Bo Γ (s),
which can be written as
or
Pre-multiplying by C = Nc yields
which is also of full row rank due to nonsingularity of the second matrix on
the right hand side. Thus, (A, B) is controllable.
Theorem 2.3.8. Let N (s)D−1 (s) be a right PMF of a strictly proper ratio-
nal function matrix G(s) with D(s) column-reduced. Then
(i) (A, B, C) in (2.44)-(2.46) is a controllable realization of G(s);
(ii) (A, B, C) is a controllable and observable (or minimal) realization of
G(s) if N (s)D−1 (s) is coprime; and
44 2. Representations of Linear Dynamic Systems
The proof of (ii) and (iii) can be found, say, in Chen (1984). With obvious
changes, dual results to Theorems 2.3.7 and 2.3.8 can be established for
properness and realization of left PMFs.
One gets
1
−1 (s + 1)(s + 2) 0 1 s+2
D̄(s) := D̃(s)R (s) = 1
0 s+2 0 s+2
(s + 1)(s + 2) s + 1
=
0 1
and
1
−1 2s + 3 1 1 s+2
N̄ (s) := Ñ (s)R (s) = 1
s+1 1 0 s+2
2s + 3 2
=
s+1 1
for a right coprime PMF N̄ (s)D̄−1 (s) of G(s).
But D̄(s) is not column reduced, since its highest column degree coefficient
matrix,
11
D̄hc = ,
00
is singular. We see that (1,1) element will become zero if the 2nd column is
multiplied by −(s+2) and the result is added to the first column. This means
2.3 Rational Matrices and Polynomial Matrix Fractions 45
(s + 1)(s + 2) s + 1 1 0
D̄(s)U (s) =
0 1 −(s + 2) 1
0 s+1
= := D(s),
−(s + 2) 1
which is now column reduced as desired. The corresponding numerator in the
right coprime PMD N (s)D−1 (s) = G(s) is
2s + 3 2 1 0
N (s) := N̄ U (s) =
s+1 1 −(s + 2) 1
−1 2
= .
−1 1
Using the realization formulas, we have
µ
s 1 0 s0
Sc (s) = = ,
0 sµ2 0s
10
Φc (s) = = I2 ,
01
0 s+1 0 1 s0 0 1
D(s) = = + I ,
−(s + 2) 1 −1 0 0s −2 1 2
−1 2
N (s) = I = Nc Φ c ,
−1 1 2
00 10
Ao = , Bo = ,
00 01
−1 −1 −2 1
A = Ao − Bo Dhc Dlc = −Dhc Dlc = ,
0 −1
−1 −1 0 −1
B= Bo Dhc = Dhc = ,
1 0
−1 2
C = Nc = .
−1 1
Reduced-order models are often required for simplifying the design and im-
plementation of control systems. A reduced-order model is usually adequate,
provided it has dynamic characteristics close to that of the original high-order
system in the frequency range which is most relevant for control system de-
sign. Therefore, model reduction has been an active research area in engineer-
ing, especially in model-based prediction, control and optimization, and is a
key step in many designs to be presented in later chapters. For plants with
rational transfer functions, balanced truncation and optimal Hankel norm
approximation are popular. However, the transfer functions which will ap-
pear in the subsequent chapters usually contain time delay and may even
not fit into the form of a rational function plus time delay. In this section,
we present two approaches to model reduction, based on the recursive Least
Squares and step response construction, respectively. A great attention is
paid to the preservation of stability for the reduced-order models. These two
approaches are generally applicable provided that plant transfer function or
frequency response is available.
where
(k) W (jωi )
W̄i , (k−1) (k−1) (k−1)
(2.51)
(jωi )n + αn−1 (jωi )n−1 + ... + α1 (jωi ) + α0
acts as a weighting function in the standard least squares problem in braces.
Equation (2.50) is re-arranged to yield
M
X
(k) (k) (k)T (k) 2
e0 , |ηi − φi θ | ,
i=1
where
(k) (k)
ηi = −g0 (jωi )(jωi )n W̄i , (2.52)
(k) (k) (k) (k) (k) (k) T
θ = [ αn−1 ... α0 βn ... β1 β0 ] , (2.53)
(k) (k)
φi = [ g0 (jωi )(jωi )n−1 . . . g0 (jωi ) −(jωi )n . . . −(jωi ) −1 ]T W̄i . (2.54)
Once the above RLS in (2.55)-(2.56) has been completed, the resultant pa-
(k)
rameter vector θ(k) = θ(k,M ) is used to update W̄i to
(k+1) 1
W̄i = (k) (k) (k)
, (2.59)
(jωi )n + αn−1 (jωi )n−1 + ... + α1 (jωi ) + α0
Simulation shows that the most important frequency range for the model
reduction is a decade above and below ωc , where ωc is the unity gain cross-
over frequency of the transfer function G0 (s). Therefore, the frequency range
[ω1 , ωM ] in the optimal fitting problem (2.49) is chosen to span M logarith-
mically equally spaced points between 0.1ωc and 10ωc .
Once a reduced-order model Ĝ(s) is found, the following frequency re-
sponse maximum relative estimation error can be evaluated
Ĝ(jω) − G(jω)
E := max | |. (2.60)
ω∈(0, ωM ) G(jω)
E≤ (2.61)
k∆L, k = 0, 1, · · · , N .
step 2. Start from Ĝ(s) with n = 1.
step 3. For each Lk , find the nth order rational approximation solution ĝ0 (s)
to the modified process g0 (s) = G(s)eLk s with the RLS method in
(2.55) - (2.56) and evaluate the corresponding approximation error
E in (2.61) for Ĝ(s) = ĝ0 (s)e−Lk s .
2.4 Model Reduction 49
step 4. Take Ĝ(s) as the solution if it yields the minimum error E and E ≤ .
Otherwise, n + 1 → n and go to Step 3.
The preservation of stability is a crucial issue in frequency domain based
model reduction. Let Ĝ(s) = ĝ0 (s)e−Ls be the one that yields the smallest
approximation error e in (2.48). It is noted that Algorithm 2.4.1 with zero ini-
tial parameter vector might result in unstable Ĝ(s), especially for high-order
models, even though G(s) is stable. One can use the stability tests and pro-
jection algorithm (Ljung and Soderstrom, 1983) to constrain the poles of the
model to be in the stable region. However, simulation shows that this method
can slow down the convergence of the recursive least square considerably and
may result in large modelling error, if the dynamics of the transfer function
to be modelled is complicated. Here, we notice that since (2.48) is a nonlin-
ear problem, Algorithm 2.4.1 may yield different local optimal solutions, if it
starts with different initial settings. Among those solutions, only stable Ĝ(s)
models are required given a stable G(s). If we initiate the algorithm with
a stable model, the algorithm is likely to reach a stable approximate upon
convergence.
When n = 1, we set the initial model for Ĝ(s) as
β0
Ĝ0 = e−L0 s . (2.62)
s + α0
Similarly, match Ĝ0 (jω) to G(jω) at the two points ω = ωb and ω = ωc , where
∠G(jωb ) = − π2 and ∠G(jωc ) = −π. It then follows (Wang et al., 1999a) that
the parameters β0 , α1 , α0 and L0 can be determined as
sin(ωc L0 ) ωc |G(jωc )|
cos(ωb L0 ) = ωb |G(jωb )| ,
h i−1
β0 = (ωc2 − ω 2 ) sin(ωb L0 ) + cos(ωc L0 )
b |G(jωb )| |G(jωc )| ,
sin(ωc L0 ) (2.65)
α1 /β0 = ωc |G(jωc )| ,
h i−1
α /β = (ω 2 − ω 2 ) ωc2 sin(ωb L0 ) + ωb2 cos(ωc L0 )
0 0 c b |G(jωb )| |G(jωc )| .
50 2. Representations of Linear Dynamic Systems
For the cases of n > 2, we can set the initial model Ĝ0 (s)of the respective
orders with β0 , α1 , α0 and L0 determined as in (2.65), while all the remaining
high-degree coefficients are set to 0. Our extensive simulation shows that this
technique works very well.
0.5
−0.5
−1
−1.5
−2
−1 −0.5 0 0.5 1 1.5 2 2.5
G(0)
Y (s) = + L{∆y(t)}. (2.69)
s
Bring (2.68) in, we have
G(s) − G(0)
L{∆y(t)} = h .
s
Applying the inverse Laplace transform yields
G(s) − G(0)
∆y(t) = hF −1 { }.
s
Thus, the plant step response is constructed as
G(jω) − G(0)
y(t) = h[G(0) + F −1 { }], (2.70)
jω
52 2. Representations of Linear Dynamic Systems
Identification from Step Response. Suppose that the given stable step
response is to be fitted into the following model:
or equivalently by
y (n) (t) + a1 y (n−1) (t) + . . . + an−1 y (1) (t) + an y(t) = b1 u(n−1) (t − L)
+b2 u(n−2) (t − L) + . . . + bn−1 u(1) (t − L) + bn u(t − L), (2.72)
where L > 0. For an integer m > 1 , define
Z (m) Z tZ τm Z τ2
f= ··· f (τ1 )dτ1 · · · dτm .
[0,t] 0 0 0
Define
2.4 Model Reduction 53
γ(t) = y(t),
h R i
(1) R (n)
φT (t) = − [0,t] y, · · · , − [0,t] y, h, ht, ht2 tn h
h 2! , · · · , n! ,
T
Pn b (−L)j Pn bj (−L)j−1 (2.74)
θ = a1 , · · · , an , j=1 j j! , j=1 (j−1)! , · · · ,
Pn i
j−(n−1)
j=n−1 bj (−L) , bn .
where Γ = [γ(t1 ), γ(t2 ), ..., γ(tN )]T , and Φ = [φ(t1 ), φ(t2 ), ...., φ(tN )]T .
In (2.75), the parameter vector θ which minimizes the following error,
min(Γ − Φθ)T (Γ − Φθ),
θ
θ̂ = (Z T Φ)−1 Z T Γ (2.81)
Once θ is estimated from (2.80), one has to recover the process model
coefficients: L, ai and bi , i = 1, 2, · · · , n. It follows from (2.74) that
ak = θk k = 1, 2, · · · , n;
and
(−L)2 (−L)3 (−L)n−1 (−L)n
−L ···
2! 3! (n−1)! n! θn+1
1 −L (−L)
2
··· (−L)n−2 (−L)n−1 b1
2! (n−2)! (n−1)! θn+2
b2
(−L)n−3 (−L)n−2 θn+3
0 1 −L · · · (n−3)! (n−2)!
b3
n−3 θn+4
0 0 1 ··· (−L) n−4
(−L)
=.. . (2.82)
(n−4)! (n−3)! ..
.
.. .. .. .. .. .. .
. . . . . . bn−1
θ2n
0 0 0 ··· 1 −L bn θ2n+1
0 0 0 ··· 0 1
(L)n−2 (L)n−1
θ
L2 n+2
b1 1 L · · ·
2! (n−2)! (n−1)! θn+3
b2 0 1 L · · · (L)n−3 (L)n−2
(n−3)! (n−2)! θn+4
..
. = .. .. .. . . ... ... · .. . (2.83)
. . . .
.
bn−1
0 0 0 ··· 1 L θ2n
bn 0 0 0 ··· 0 1 θ2n+1
Substituting (2.83) into the first row of (2.82) yields the following n-degree
polynomial equation in L:
n
X θn+i+1
Li = 0. (2.84)
i=0
i!
with
( p
− θ2 − 2θ5 θ3 , if a inverse response is detected;
β= p 24
θ4 − 2θ5 θ3 , otherwise.
where y(ti ) and ŷ(ti ) are the step responses of the plant G(s) and its model
Ĝ(s), respectively, under the same step input. Once the identification is car-
ried out, the model Ĝ(s) is available and the frequency error E in (2.60) is
measured as the worst-case error. Here, the frequency range [0, ωc ] is con-
sidered, where ∠G(jωc ) = −π, since this range is the most significant for
control design. To show the robustness of the proposed method, noise is in-
troduced into the output response. In the context of system identification,
noise-to-signal ratio defined by
mean(abs(noise))
N SR = .
mean(abs(signal))
is used to represent noise level.
2.5 Conversions between Continuous and Discrete Systems 57
Example 2.4.2. Reconsider the high-order plant in Example 2.4.1. The out-
put step response is first constructed using the IFFT. The area method
(Rake, 1980) gives the model:
2.15
Ĝ(s) = e−53.90s ,
46.69s + 1
with = 1.15 × 10−2 and E = 60.87%. The FOPDT model estimated by
Algorithm 2.4.2 is
0.0396 −49.9839s
Ĝ(s) = e ,
s + 0.0184
with = 8.6 × 10−3 and E = 48.12%. Our SOPDT model is
0.0011
Ĝ(s) = e−28.8861s ,
s2 + 0.0343s + 0.0005
with = 4.0631 × 10−4 and E = 5.81%. For third-order plus dead time model
(TOPDT), we have
This book will address control system design for continuous-time systems
only. For computer implementation, the designed continuous controllers have
to be converted to discrete forms. In general, signals from computer related
applications are often manipulated in digital form. Conversion between con-
tinuous and discrete time systems is thus needed in these applications. The
conversion accuracy is important and directly affects the final performance of
58 2. Representations of Linear Dynamic Systems
2.5
1.5
0.5
−0.5
0 50 100 150 200 250 300 350
0.5
−0.5
−1
−1.5
−2
−1.5 −1 −0.5 0 0.5 1 1.5 2 2.5
Unite circle
Fig. 2.3. Maps of the left-half s-plane to the z-plane by the integration rules in
Table .2.1. Stable s-plane poles map into the shaded regions in the z-plane. The
unite circle is shown for reference. (a) Forward rectangular rule. (b) Backward
rectangular rule.(c) Trapezoid or bilinear rule.
G(s)|s=0 = G(z)|z=1 .
(z + 1)(1 − e−aT )
G(z) = .
2(z − e−aT )
1 − e−aT
G(z) = .
z − e−aT
Hold Equivalents. In this case, we want to design a discrete system G(z)
that, with an input consisting of samples of a continuous signal u(t), has an
output which approximates the output of the continuous system G(z) whose
input is the continuous signal u(t). We can do so through hold operation. If
the hold is of the zero-order, the conversion is given by
G(s)
G(z) = (1 − z −1 )Z{ }.
s
Suppose again
a
G(s) = .
s+a
2.5 Conversions between Continuous and Discrete Systems 61
It follows that
G(s) a 1 1
Z{ } = Z{ } = Z{ } − Z{ }
s s(s + a) s s+a
(1 − e−aT z −1 ) − (1 − z −1 )
= ,
(1 − z −1 )(1 − e−aT z −1 )
so that
1 − e−aT
G(z) = .
z − e−aT
The rules discussed above are applicable mainly to rational G(s) only. We
now develop a conversion technique which is suitable for non-rational G(s),
too. Let a continuous transfer function Gc (s) or frequency response Gc (jω)
be given. An equivalent discrete transfer function
is to be found to match the frequency response of Ĝd (z) to Gc (s) with respect
to the map,
z = eT s . (2.87)
where (ω1 , ωM ) defines the frequency range of interest. Equation (2.88) falls
into the framework of transfer function identification in frequency domain, in
which Gc is the known frequency response while Ĝd is a parametric transfer
function to be identified. Obviously, the RLS based method (Algorithm 2.4.1)
in the preceding section is immediately applicable. To make the best use of
it, two minor remarks are given as follows.
The frequency range in the optimal fitting is a useful tuning parameter.
In general, our studies suggest that the frequency range (ω1 · · · ωM ) be cho-
1 1 1
sen as ( 100 ωb , ωb ) with the step of ( 100 ∼ 10 )ωb , where ωb is the bandwidth
of Gc (jω). In this range, RLS yields satisfactory fitting results in frequency
domain. Next, the preservation of stability is again a crucial issue. We adopt
the same strategy, i.e., make the initial model stable. But, the formulas in
(2.62) and (2.64) are no longer useful as they only ensure stability of contin-
uous systems. To get a stable initial Gd0 (z) from the given stable Gc (s), note
from Figure 2.3(a) that the Bilinear rule preserves stability under continuous
62 2. Representations of Linear Dynamic Systems
and discrete conversions. Thus, for a rational Gc (s), we set the initial model
Gd0 (z) according to the Bilinear rule. For other Gc (s), the idea behind (2.62)
and (2.64) can still be used, that is, match Gd0 (z) to Gc (s) at two selected
frequencies to get a stable Gd0 (z). Our extensive simulation shows that this
technique works very well.
The method described above is also applicable to inverse conversion, i.e.,
from DTF to CTF, with obvious changes of the rules of continuous and
discrete frequency responses in (2.88).
We now illustrate the above technique with some typical plants. In sim-
ulation, the sampling time T is chosen to meet the standard rule:
ωs
6. . 40, (2.89)
ωb
where ωs = 2πT is the sampling frequency. Once a discrete equivalent is found,
the frequency response maximum relative error similar to (2.60) is used to
assess the conversion accuracy.
Example 2.5.1. Consider the continuous filter transfer function (Franklin et
al., 1990, pp. 141):
1
Gc (s) = ,
s3 + 2s2+ 2s + 1
ωs
with unity pass bandwidth. For T = 1 sec, i.e., ωb = 2π > 6, the Bilinear
rule gives
0.0476z 3 + 0.1429z 2 + 0.1429z + 0.0476
Ĝd (z) = ,
z 3 − 1.1900z 2 + 0.7143z − 0.1429
with the error E = 24.67 %, while the zero-pole mapping transformation does
0.0920(z + 1)2
Ĝd (z) = ,
(z − 0.3679)(z 2 − 0.7859z + 0.3679)
with E = 62.84 %. For m = 3, the fitting method yields
0.0080z 3 + 0.2257z 2 + 0.1381z − 0.0072
Ĝd (z) = ,
z 3 − 1.1583z 2 + 0.6597z − 0.1367
with E = 0.26 %. In Figure 2.4, the responses of three models are compared,
and the fitting method yields much better performance than others. More-
over, Figure 2.5 shows that in the reasonable range of ωs given in (2.89),
the fitting method exhibits a consistent superior approximation performance
than Bilinear and zero-pole mapping schemes. ♦
Example 2.5.2. Consider the system with time delay
s−1
Gc (s) = e−0.35s , (2.90)
s2 + 4s + 5
2.5 Conversions between Continuous and Discrete Systems 63
0.8
0.6
gain
0.4
0.2
−50
−100
phase (°)
−150
−200
−250
−300
with T = 0.1 sec (Control System Toolbox User’s Guide, MathWorks, ver. 4.2,
1998, c2d function). The Toolbox, via Triangle approximation, gives
To end this section, it is worth mentioning that the fitting method is suit-
able for general systems with either parametric or non-parametric source
models, and can achieve high performance with good chance of stability
preservation. In fact, it can be also applied to many other cases such as
determining optimal sampling time, optimal model order, and re-sampling a
discrete systems. Readers are referred to Wang and Yang (2001) for details.
64 2. Representations of Linear Dynamic Systems
0
10
E
−1
10
−2
10
−3
10
−4
10
−5
10
−6
10
10 15 20 25 30 35 40
ws/wb
Feedback is the main tool for control systems. Stability and robustness are
two key issues associated with any feedback design, and are the topics of this
chapter. Section 1 addresses internal stability of general interconnected sys-
tems and derives a powerful stability condition which is applicable to systems
with any feedback and/or feedforward combinations. Section 2 focuses on the
conventional unity output feedback configuration and gives the simplifying
conditions and Nyquest-like criteria. The plant uncertainties and feedback
system robustness to them are discussed in Section 3, both structured and
unstructured perturbations are introduced. A special attention is paid to the
case where the phase perturbation is limited, a realistic situation.
A−1
i (s)Bi (s)ÿ is a left coprime polynomial matrix fraction (PMF). Then
pi (s) = det Ai (s) is the characteristic polynomial of Gi (s). Gi (s) can be
considered (Rosenbrock 1974) to arise from the equations:
Ai (s)ξi (s) = Bi (s)ui (s), (3.1a)
yi (s) = ξi (s), (3.1b)
where ui and yi are the input and output vectors of plant Gi (s) respectively.
n plants in form of (3.1) may be combined to
A(s)ξ(s) = B(s)U (s), (3.2a)
Y (s) = ξ(s), (3.2b)
where
A(s) = block diag{A1 (s), A2 (s), · · · , An (s)},
B(s) = block diag{B1 (s), B2 (s), · · · , Bn (s)},
T
ξ = ξ1T ξ2T · · · ξnT ,
T
U = uT1 uT2 · · · uTn .
In an interconnected system, the outputs of some plants may be injected into
other plants as inputs. Besides, various exogenous signals (references and
disturbances) rj , j = 1, 2, · · · , m, may also come into the system. Thus, U
has the general form:
has all its roots in the open left half of the complex plane.
for proper G(s). Obviously, Assumption 3.1.1 is true if G(s) is strictly proper,
i.e., all plants Gi (s) are strictly proper.
though G(s) = 1−s s is proper. For this case, one sees that U = R − Y and
C = −1, which leads to 1 − G(s)C = 1 + G = 1s and det(1 − G(∞)C) = 0,
violating Assumption 3.1.1.
Recall that a rational function is stable if it is proper and all its poles have
negative real parts. The properness has to be included, otherwise, a bounded
input may generate an unbounded output even though all the poles of the
rational function have negative real parts. Take the above example again for
instance, one has pc (s) = det(A(s) − B(s)C) = (s + 1 − s) = 1 for which there
is no pole and thus no unstable pole. But HY R = 1 − s, for the bounded
input, say, r(t) = sin(t2 ), the output response under zero initial condition is
y(t) = sin(t2 ) − 2t cos(t2 ) and unbounded.
68 3. Stability and Robustness of Feedback Systems
Refer back to the interconnected system described in (3.2) and (3.3), for
ease of future reference, we define the system connection matrix as
The system which Theorem 3.1.1 can be applied to can contain multi-
input plants. In such a case, it is here called the system with vector signals
since its input and/or output are vector signals. On the other hand, if a lin-
ear time-invariant interconnected system consists only of SISO plants Gi (s),
then such a system is here called the system with scalar signals because all
the signals in the system is scalar. For such systems, we are going to de-
velop an elegant stability criterion which makes use of plants’ characteristic
polynomials and the system determinant only without evaluating det W (s)
(which may be difficult to calculate for large-scale systems). This criterion
may also be applied to systems with vector signals if all multi-input and/or
multi-output plants involved in the system are viewed to consist of several
SISO subsystems and are represented by their coprime polynomial matrix
fractions (PMFs) or minimal state-space realizations in the signal flow graph
of the system.
For a system with scalar signals, we first draw its signal flow graph. Recall
that a signal flow graph consists of nodes and directed branches. A node per-
forms the functions of addition of all incoming signals and then transmission
of it to all outgoing branches. A branch connected between two nodes acts
as a one-way signal multiplier and the multiplication factor is the transfer
function of the corresponding plant. A loop is a closed path that starts at a
node and ends at the same node. We say that a signal goes through a loop
if and only if it passes through all nodes and branches in the loop once. If
two loops have neither common nodes nor branches, then they are said to be
nontouched. The loop gain for a loop is the product of all transfer functions
of the plants in the loop. The system determinant ∆ is defined (Mason 1956)
by
3.1 Internal Stability 69
X X X
∆=1− F1i + F2j − F3k + · · · , (3.8)
i j k
where F1i are the loop gains, F2j are the products of 2 non-touching loop
gains, F3k are products of 3 non-touching loop gains, · · · . The following is
the major result of this section.
Theorem 3.1.2. Let an interconnected system consist of n SISO plants
Gi (s), i = 1, 2, · · · , n, and the characteristic polynomial of Gi (s) be pi (s)
(simply the transfer function denominator) for each i. Suppose ∆(∞) 6= 0.
Then, the system is internally stable if and only if
n
Y
pc (s) = ∆(s) pi (s) (3.9)
i=1
has all its roots in the open left half of the complex plane.
Some insights may be drawn from the above theorem. Firstly, any pole-
zero cancellations will be revealed by pc (s) in (3.9). To see this, let p(s) be a
least commonQdenominator of all the loop gains after pole-zero cancellations
and po (s) = i pi (s). Then p(s) divides po (s) and the quotient contains the
poles of some plants that have been cancelled by other plant zeros but still
appear in pc (s) as the closed-loop poles. Secondly, one may see that any
unstable poles in the plants outside of the loop will be retained as the roots
of pc (s).
It should be pointed out that Theorem 3.1.1 can actually be applied to
systems with scalar signals too, since the scalar signal is a special case of a
vector signal. However, Theorem 3.1.2 is recommended in the case of scalar
signals as it is much simpler to use than Theorem 3.1.1. The internal stability
of several typical control schemes is now investigated with the help of these
results.
r + u1 y1 u2 y2 un yn
G1 G2 Gn
_
Example 3.1.1. Consider the single loop feedback system shown in Figure
3.2. It follows that
70 3. Stability and Robustness of Feedback Systems
0 0 0 ··· −I
I 0 0 ··· 0
0
W = I − GC = I − diag{Gi } 0 I 0 ···
.. .. .. . . ..
. . . . .
0 0 0 ··· 0
I 0 0 · · · G1
−G2 I 0 ··· 0
··· 0
= 0 −G3 I .
.. .. .. . . ..
. . . . .
0 0 0 ··· I
Post-multiplying the i-th column by Gi Gi−1 · · · G1 and subtracting it from
the n-th column, i = 1, 2, · · · , n − 1, yields
I 0 0 ··· 0
−G2 I 0 · · · G2 G1
det(W ) = det 0 −G3 I · · · 0
.. .. .. . . ..
. . . . .
0 0 0 ··· I
= ···
= det[I + Gn Gn−1 · · · G1 ].
Thus, it follows from Theorem 3.1.1 that Qnthe system is internally stable if and
only if pc (s) = det(I + Gn Gn−1 · · · G1 ) i=1 pi (s) has all its roots in the open
left half of the complex plane, where pi (s) are the characteristic polynomials
of Gi (s) respectively. ♦
Example 3.1.2. Consider the Smith predictor control system depicted in Fig-
ure 3.3. Assume that all the plants in the system are of SISO. Let g(s) =
g0 (s)e−Ls , g0 (s) = b(s)/a(s) and k(s) = β(s)/α(s) and be coprime polyno-
mial fractions. The system consists of 4 plants, g(s), g0 (s), e−Ls and k(s).
Their characteristic polynomials are a(s), a(s), 1 and α(s) respectively, so
Q4
that i=1 pi (s) = a2 (s)α(s). To find the system determinant ∆(s), one easily
3.1 Internal Stability 71
r + y
k (s ) g (s )
_
_ +
g 0 (s) e - Ls
+
+
sees that the system has three loops which pass through (k, g0 ), (k, g0 , e−Ls )
and (k, g) respectively. Note that all these loops have the common branch k(s)
and thus they are touched. It follows that ∆ = 1+kg0 +kg−kg0 e−Ls = 1+kg0 .
According to (3.9), pc (s) for this Smith system is
4
Y ÿ
pc (s) = ∆(s) pi (s) = b(s)β(s) + a(s)α(s) a(s).
i=1
ÿ
It is well known that b(s)β(s) + a(s)α(s) can be assigned to an arbitrary
stable polynomial by a proper choice of α(s) and β(s) in k(s) for coprime
a(s) and b(s). Hence, the system is internally stabilizable if and only if a(s)
is a stable polynomial, or if and only if the process g(s) is stable. ♦
Example 3.1.3. Consider the input-output feedback system shown in Figure
3.4. Let g(s) = b(s)/a(s) be a coprime polynomial fraction, pf (s) and q(s) be
two given polynomials, and f (s) = a(s) − pf (s). Polynomials k(s) and h(s)
are determined by solving the polynomial equation:
k(s)a(s) + h(s)b(s) = q(s)f (s).
To analyze the system with our method, note that there are two touched
loops in Figure 3.4 whose gains are k(s)/q(s) and g(s)h(s)/q(s). It follows
that
k(s) h(s)
pc (s) = 1 − − g(s) q(a)a(s)
q(s) q(s)
= q(s)a(s) − k(s)a(s) − h(s)b(s) = q(s)pf (s),
where both q(s) and pf (s) are specified by the user. Thus, all the poles of
the system can be assigned arbitrarily. ♦
v + u y
g(s)
+
+ +
k(s) h(s)
1
q(s)
g7 g4
g1 g2
g3
g5
g6
involves matrix calculations while ∆ deals with the system signal flow graph.
To link them together, we will perform a series of expansions on det(W ) and
express det(W ) in terms of those items which have clear meanings on the
signal flow graph. It should be pointed out that the method we will use to
prove Theorem 3.1.2 below is general and applicable to any interconnected
system. But for ease of understanding, we shall frequently refer to an example
for demonstration when we introduce notations, concepts, derivations and
results. The example is the signal flow graph in Figure 3.5, and according to
(3.11) its system connection matrix W easily reads as
1 0 0 0 0 −g1 −g1
−g2 1 0 0 −g2 0 0
0 −g3 1 −g3 0 0 0
W = 0 0 −g4 1 0 0 0 . (3.12)
0 −g5 0 −g5 1 0 0
0 0 −g6 0 0 1 0
−g7 0 0 0 −g7 0 1
Let Wi1 ,i2 ,··· ,iv denote the system connection matrix of the resultant sys-
tem obtained by removing plants gi1 , gi2 , · · · , giv from the system. For each
gik , k = 1, 2, · · · , v, the removal of gik from the system is equivalent to the
system where there is no signal out of gik and no signal injected into it.
Then it is obvious that matrix Wi1 ,i2 ,··· ,iv is identical to the matrix formed
by deleting columns (i1 , i2 , · · · , iv ) and rows (i1 , i2 , · · · , iv ) from W . For ex-
ample, if plants g3 and g5 are removed from the system in Figure 3.5, then its
signal flow graph becomes Figure 3.6. It follows from (3.11) that the system
connection matrix is
1 0 0 −g1 −g1
−g2 1 0 0 0
W3,5 = 0 0 1 0 0 . (3.13)
0 00 1 0
−g7 0 0 0 1
g7 g4
g1 g2
g6
Fig. 3.6. Signal Flow Graph by deleting Plants g3 and g5 in Figure 3.5
74 3. Stability and Robustness of Feedback Systems
Indeed, (3.13) is identical to the matrix in (3.12) after deleting the third and
fifth columns and the third and fifth rows. Recall (Griffel, 1989) that a minor
of a matrix is the determinant of its submatrix. Let Mji11,j ,i2 ,··· ,iv
2 ,··· ,jv
be the minor
of W formed by deleting rows (i1 , i2 , · · · , iv ) and columns (j1 , j2 , · · · , jv ) from
W . We have just shown that matrix Wi1 ,i2 ,··· ,iv is identical to the matrix
formed by deleting columns (i1 , i2 , · · · , iv ) and rows (i1 , i2 , · · · , iv ) from W .
Thus, the following fact is true.
Fact 1 det(Wi1 ,i2 ,··· ,iv ) = Mii11,i,i22,···i
,··· ,iv
v
.
The key idea to prove Theorem 3.1.2 is to expand det(W ) step by step and
express it in terms of smaller sizes of minors until those minors are reached
which correspond to some closed-paths in the system signal graph. For later
reference, we define a closed-path as a signal flow path which starts from
a plant, may pass through other plants and nodes, and comes back to the
same plant. It should be noted that a closed-path may pass through a node
many times but a loop (Mason, 1956) is a special closed-path which can
pass through a node only once. For example, g1 , g2 , g3 and g6 form a loop,
and g1 , g2 , g5 and g7 form a closed-path but not a loop. Turn back to the
expansion of det(W ), and use W in (3.12) as an example. At the first stage,
we expand det(W ) by the first column as
ü ü
ü 1 0 0 0 0 −g1 −g1 üü
ü
ü −g2 1 0 0 −g2 0 0 üü
ü
ü 0 −g3 1 −g3 0 0 0 üü
ü
det(W ) = üü 0 0 −g4 1 0 0 0 üü
ü 0 −g5 0 −g5 1 0 0 ü
ü ü
ü 0 0 −g 0 0 1 0 ü
ü 6 ü
ü −g7 0 0 0 −g7 0 1 ü
ü ü
ü 0 0 0 0 −g1 −g1 üü
ü
ü −g3 1 −g3 0 0 0 üü
ü
ü 0 −g 1 0 0 0 üü
= det(W1 ) + (−1)4 × g2 üü 4
ü
ü −g 5 0 −g 5 1 0 0 ü
ü 0 −g6 0 0 1 0 ü
ü ü
ü 0 0 0 −g7 0 1 ü
ü ü
ü 0 0 0 0 −g1 −g1 üü
ü
ü 1 0 0 −g2 0 0 üü
ü
ü −g3 1 −g3 0 0 0 üü
ü
9 ü
+(−1) × g7 ü 0 −g4 1 0 0 0 üü
ü −g5 0 −g5 1 0 0 üü
ü
ü 0 −g6 0 0 1 0 üü
ü
ü ü
= det(W1 ) + (−1)4 g2 M12 + (−1)9 g7 M17 . (3.14)
One may note from (3.14) that det(W ) has two items left in addition to
det(W1 ). This is because the output of g1 is injected only into g2 and g7 . The
3.1 Internal Stability 75
second stage of the expansion then deals with those nonzero minors (M12 and
M17 ) left from the first stage expansion. The columns on which we expand
these minors further are specific and are chosen as follows. For the minor M1i ,
we expand it on the i-th column of W (note that it is the i-th column of W
but not of M1i ). For M12 , this rule leads to
ü ü
ü 0 0 0 0 −g1 −g1 üü
ü
ü −g3 1 −g3 0 0 0 üü
ü
ü 0 −g 1 0 0 0 üü
M12 = üü 4
ü
−g
ü 5 0 −g 5 1 0 0 ü
ü 0 −g6 0 0 1 0 ü
ü ü
ü 0 0 0 −g7 0 1 ü
ü ü ü ü
ü 0 0 0 −g1 −g1 üü ü 0 0 0 −g1 −g1 üü
ü ü
ü −g4 1 0 0 0 üü ü 1 −g3 0 0 0 üü
ü ü
= (−1)4 g3 üü 0 −g5 1 0 0 üü + (−1)6 g5 üü −g4 1 0 0 0 üü
ü −g6 0 0 1 0 ü ü −g6 0 0 1 0 ü
ü ü ü ü
ü 0 0 −g7 0 1 ü ü 0 0 −g7 0 1 ü
2,3 2,5
= (−1)4 g3 M1,2 + (−1)6 g5 M1,2 , (3.15)
2,3 2,5
from which one may note once again that M1,2 and M1,2 are left because the
output of g2 is injected into g3 and g5 . Similarly, we expand M17 on the 7-th
column of W as
ü ü
ü 0 0 0 0 −g1 −g1 üü
ü
ü 1 0 0 −g2 0 0 üü
ü
ü −g 1 −g 0 0 0 üü
M17 = üü 3 3
ü 0 −g 4 1 0 0 0 üü
ü −g5 0 −g5 1 0 0 ü
ü ü
ü 0 −g6 0 0 1 0 ü
ü ü
ü 1 0 0 −g2 0 üü
ü
ü −g3 1 −g3 0 0 ü
ü ü
= (−1)8 g1 üü 0 −g4 1 0 0 üü = (−1)8 g1 M1,7
7,1
. (3.16)
ü −g5 0 −g5 1 0 ü
ü ü
ü 0 −g6 0 0 1ü
Note from Figure 3.5 that the output of g7 comes back to the input of g1
7,1
and form a closed-path. The corresponding minor M1,7 , by Fact 1, is equal
to det(W1,7 ), the determinant of the system connection matrix formed by
removing all plants in this closed-path. In general, if the plants involved
in a particular minor Mji11,j ,i2 ,··· ,iv
2 ,··· ,jv
form a closed-path, i.e., {i1 , i2 , · · · , iv } =
{j1 , j2 , · · · , jv } (namely, two sets are equal), we stop expanding this minor
and express it as det(Wi1 ,i2 ,··· ,iv ); Otherwise, continue to expand the minor.
With this rule, we proceed from (3.15) as
76 3. Stability and Robustness of Feedback Systems
ü ü
ü 0 0 0 −g1 −g1 üü
ü
ü −g41 0 0 0 üü
ü
2,3
M1,2 = üü 0 −g5 1 0 0 üü
ü −g60 0 1 0 üü
ü
ü 0 0 −g7 0 1 ü
ü ü ü ü
ü 0 0 −g1 −g1 üü ü 0 0 −g1 −g1 üü
ü ü
4 ü
ü −g 1 0 0 ü ü
6 ü 1 0 0 0 üü
= (−1) g4 ü 5 ü
0 0 1 0 ü + (−1) g6 ü −g5 1 0 0 üü
ü ü ü
ü 0 −g7 0 1 ü ü 0 −g7 0 1 ü
2,3,4 2,3,6
= (−1)4 g4 M1,2,3 + (−1)6 g6 M1,2,3 , (3.17)
and
ü ü
ü 0 0 0 −g1 −g1 üü ü ü
ü ü 0 0 −g1 −g1 üü
ü 1 −g3 0 0 0 ü ü ü
ü ü
9 ü 1 −g3 0 0 üü
2,5
M1,2 = üü −g4 1 0 0 ü
0 ü = (−1) g7 ü
ü −g6 ü −g4 1 0 0 üü
ü 0 0 1 0 üü ü −g6
ü 0 0 1 0 ü
0 −g7 0 1 ü
2,5,7
= (−1)9 g7 M1,2,5 . (3.18)
2,5,7 2,3,4 2,3,6
Since the minors M1,2,5 , M1,2,3 and M1,2,3 do not satisfy the condition on
stopping expansions, we continue to expand them as
ü ü
ü 0 0 −g1 −g1 üü ü ü
ü ü 1 −g3 0 ü
ü 1 −g3 0 ü
0 ü ü ü
2,5,7
M1,2,5 = üü ü = (−1)6 g1 üü −g4 1 0 üü
ü −g4 1 0 0 ü ü −g6 0 1 ü
ü −g6 0 1 0 ü
2,5,7,1
= (−1)6 g1 M1,2,5,7 = (−1)6 g1 det(W1,2,5,7 ), (3.19)
where it has been noticed that plant g1 , g2 , g5 and g7 form a closed-path,
which is clear from Figure 3.5 and also from the fact {1, 2, 5, 7} = {2, 5, 7, 1},
and no further expansion on W1,2,5,7 will be carried out. Similarly,
ÿ ÿ
ÿ 0 −g1 −g1 ÿ
ÿ ÿ
2,3,4
M1,2,3 = (−1) g5 ÿÿ 0
4
1 0 ÿÿ = (−1)4 g5 M1,2,3,4
2,3,4,5
,
ÿ −g7 0 1 ÿ
ÿ ÿ
ÿ −g −g1 ÿ
= (−1)4 g5 (−1)5 g7 ÿÿ 1 ÿ = (−1)9 g5 g7 M1,2,3,4,5
2,3,4,5,7
1 0 ÿ
= (−1)9 g5 g7 (−1)4 g1 M1,2,3,4,5,7
2,3,4,5,7,1
= (−1)13 g5 g7 g1 det(W1,2,3,4,5,7 ); (3.20)
ÿ ÿ
ÿ 1 0 0ÿ ÿ
ÿ
2,3,6
M1,2,3 = (−1)5 g1 ÿÿ −g5 1 0 ÿÿ = (−1)5 g1 M1,2,3,6
2,3,6,1
= (−1)5 g1 det(W1,2,3,6 ), (3.21)
ÿ 0 −g7 1 ÿ
where all the minors left have reached the corresponding closed-paths. To
be more clear, we substitute expressions (3.21), (3.20), (3.19), (3.18), (3.17),
(3.16), (3.15) into (3.14), and the expanding procedure looks as
3.1 Internal Stability 77
det(W )
= det(W1 ) + (−1)4 g2 M12 + (−1)9 M17
2,3 2,5 7,1
= det(W1 ) + (−1)4 g2 [(−1)4 g3 M1,2 + (−1)6 g5 M1,2 ] + (−1)9 g7 (−1)8 g1 M1,7
2,3,4 2,3,6
= det(W1 ) + (−1)4 g2 {(−1)4 g3 [(−1)4 g4 M1,2,3 + (−1)6 g6 M1,2,3 ]
2,5,7
+(−1)6 g5 (−1)9 g7 M1,2,5 } + (−1)g1 g7 det(W1,7 )
2,3,4,5
= det(W1 ) + (−1)4 g2 {(−1)4 g3 [(−1)4 g4 (−1)4 g5 M1,2,3,4
2,3,6,1 2,5,7,1
+(−1)6 g6 (−1)5 g1 M1,2,3,6 ] + (−1)6 g5 (−1)9 g7 (−1)6 g1 M1,2,5,7 }
+(−1)g1 g7 det(W1,7 )
2,3,4,5,7
= det(W1 ) + (−1)4 g2 {(−1)4 g3 [(−1)4 g4 (−1)4 g5 (−1)5 g7 M1,2,3,4,5
+(−1)6 g6 (−1)5 g1 det(W1,2,3,6 )] + (−1)6 g5 (−1)9 g7 (−1)6 g1 det(W1,2,5,7 )}
+(−1)g1 g7 det(W1,7 )
= det(W1 )
+(−1)4 g2 {(−1)4 g3 [(−1)4 g4 (−1)4 g5 (−1)5 g7 (−1)4 g1 det(W1,2,3,4,5,7 )
+(−1)6 g6 (−1)5 g1 det(W1,2,3,6 )] + (−1)6 g5 (−1)9 g7 (−1)6 g1 det(W1,2,5,7 )}
+(−1)g1 g7 det(W1,7 )
= det(W1 )
+(−1)5 g2 g3 g4 g5 g7 g1 det(W1,2,3,4,5,7 ) + (−1)3 g2 g3 g6 g1 det(W1,2,3,6 )
+(−1)3 g2 g5 g7 g1 det(W1,2,5,7 ) + (−1)g7 g1 det(W1,7 ).
X i
det(W ) = det(W1 ) + (−1)pα1 +1 giα1 M1α1 , (3.23)
iα1
i
where the output of plant g1 is an input to plant giα1 . Next, we expand M1α1
by the iα1 -th column in W as
i
X iα1 ,iα2
M1α1 = (−1)pα2 +1 giα2 M1,i α
,
1
iα2
where the output of plant giα1 is an input to plant giα2 . If iα2 = 1, i.e., plant
α1 α2 i ,i
giα1 and plant g1 form a closed-path, then stop expanding M1,i α1
and we
have
i
α1 ,iα2 1,i
M1,i α
= M1,iαα1 = det(W1,iα1 ).
1 1
iα1 ,iα2
X i ,iα2 ,iα3
M1,i = (−1)pα3 +1 giα3 M1,i
α1
,
α1 α 1 ,iα2
iα3
where the output of plant giα2 is one of the inputs of plant giα3 . Again, if
iα3 = 1, i.e., plants giα1 , giα2 andg1 form a closed-path, then stop expanding
and we have
i
α1 ,iα2 ,iα3 1,i ,i
M1,i = M1,iαα1,iαα2 = det(W1,iα1 ,iα2 ),
α 1 ,iα2 1 2
det(W )
X iα1
= det(W1 ) + (−1)pα1 +1 giα1 M1
iα1
X X iα ,iα2
= det(W1 ) + (−1)pα1 +1 giα1 (−1)pα2 +1 giα2 M1,i1α
1
iα1 iα2
X
= det(W1 ) + (−1)pα1 +pα2 +2 Fδi det(Wδi )
iα2 =1,vδ =2
i
X pα1 +1
X iα ,iα2
+ (−1) giα1 (−1)pα2 +1 giα2 M1,i1α
1
iα1 iα2 6=1
X
= det(W1 ) + (−1)pα1 +pα2 +2 Fδi det(Wδi )
iα2 =1,vδ =2
i
X pα1 +1
X X iα ,iα2 ,iα3
+ (−1) giα1 (−1)pα2 +1 giα2 (−1)pα3 +1 giα3 M1,i1α
1 ,iα2
iα1 iα2 6=1 iα3
X
= det(W1 ) + (−1)pα1 +pα2 +2 Fδi det(Wδi )
iα2 =1,vδ =2
i
X
+ (−1)pα1 +pα2 +pα3 +3 Fδi det(Wδi )
iα3 =1,vδ =3
i
X X X iα ,iα2 ,iα3
+ (−1)pα1 +1 giα1 (−1)pα2 +1 giα2 (−1)pα3 +1 giα3 M1,i1α
1 ,iα2
iα1 iα2 6=1 iα3 6=1
···
X
= det(W1 ) + (−1)pα1 +pα2 +2 Fδi det(Wδi )
iα2 =1,vδ =2
i
X
+ (−1)pα1 +pα2 +pα3 +3 Fδi det(Wδi )
iα3 =1,vδ =3
i
X
+··· + (−1)pα1 +pα2 +···+pαx +x Fδi det(Wδi )
iαv̄ =1,vδ =v̄
i
X
= det(W1 ) + (−1)pα1 +pα2 +···+pαv +v Fδi det(W1,iα1 ,··· ,iαv−1 ),
vδ
i
g4
g2
g3
g5
g6
¯ρ1
Fig. 3.7. Signal Flow Graph for ∆
ρi in case that the involved individual plants in the loop are clear from the
context. For a loop ρi , let ∆¯ρi denote the system determinant for the system
formed by removing loop ρi . Here, it should be pointed out that removing
a loop from the system means that not only all the plants but also all the
nodes in the loop are removed from the system. Let ∆ρi denote the system
determinant for the system formed by removing all the plants in loop ρi but
retaining all the relevant nodes. For example, ρ1 = ρ(1, 7) is a loop of the
system shown in Figure 3.5, then the graph formed by removing ρ1 is shown
in Figure 3.7 and the graph by removing all the plants in ρ1 only is shown
in Figure 3.8. Furthermore, ∆¯ρ1 = 1 − g3 g4 and ∆ρ1 = 1 − g2 g5 − g3 g4 . In
general, ∆¯ρi 6= ∆ρi .
g7 g4
g1 g2 g3
g5
g6
X X
∆ = 1− F1i + F2j = 1 − g1 g7 − g2 g5 − g3 g4 − g1 g2 g3 g6 + g1 g7 g3 g4
i j
= (1 − g2 g5 − g3 g4 ) − g1 g2 g3 g6 − g1 g7 + g1 g7 g3 g4 . (3.24)
The signal flow graph after removing g1 is shown in Figure 3.9 and its system
determinant is ∆1 = 1 − g3 g4 − g2 g5 , which is, of course, independent of g1 .
The remaining part which is related to plant g1 is
∆ − ∆1 = −g1 g2 g3 g6 − g1 g7 + g1 g7 g3 g4 = −g1 Z
g7 g4
g2
g3
g5
g6
P
Lemma 3.1.2. ∆ = ∆1 − Fρi (1,··· ) ∆¯ρi (1,··· ) , where ρi (1, · · · ) are the
i
loops each of which contains plant g1 .
Now, we consider the relationship between ∆ρi and ∆¯ρi . Assume that loop
ρj has a common node with ρi , then Fρj will not appear in ∆¯ρi . It is because
this node has been removed from the graph when we calculate ∆¯ρi and the
relevant plants involved in the original ρj does not form a loop in the new
system anymore. However, Fρj will appear in ∆ρi because only the plants
involved in ρi are removed from the original graph when we calculate ∆ρi ,
but the relevant nodes are retained so that ρj is still a loop in the resultant
graph. For instance, consider the system in Figure 3.5. For ρi = ρ(1, 7), we
have ∆ρi = 1 − g2 g5 − g3 g4 and ∆¯ρi = 1 − g3 g4 . The difference between them
is g2 g5 , the gain of the ρj = ρ(2, 5) which is touched with loop ρi only by
82 3. Stability and Robustness of Feedback Systems
one node. It follows that ∆¯ρi − ∆ρi = g2 g5 = Fρ(2,5) . One further sees that
Fρi (∆¯ρi − ∆ρi ) = Fρi Fρj is the gain of the closed-path which is formed by
ρ(1, 7) and ρ(2, 5) that have a common node. For a general system, let ρij
(ρi,j
k , · · · , respectively) be the loops each of which has a common node with
ρi(1,··· ) (ρi or ρj ,· · · , respectively) but does not contain the plant g1 . It follows
from Lemma 3.1.2 that
X
∆ = ∆1 − Fρ (1,··· ) ∆¯ρ (1,··· )
i i
i
X
= ∆1 − Fρi (1,··· ) (∆ρi (1,··· ) + ∆¯ρi (1,··· ) − ∆ρi (1,··· ) )
i
X X
= ∆1 − Fρi (1,··· ) ∆ρi (1,··· ) − Fρi (1,··· ) (∆¯ρi (1,··· ) − ∆ρi (1,··· ) )
i i
X X X
= ∆1 − Fρi (1,··· ) ∆ρi (1,··· ) − Fρi [ Fρij ∆¯ρi ρij
i i j
X
− Fρij Fρij ∆¯ρi ρij ρij + ···]
1 2 1 2
j1 ,j2
X X X
= ∆1 − Fρi (1,··· ) ∆ρi (1,··· ) − Fρi Fρij ∆ρi ρij
i i j
X X X
− Fρi [ Fρij (∆¯ρi ρij − ∆ρi ρij ) − Fρij Fρij ∆¯ρi ρij ρij + · · ·]
1 2 1 2
i j j1 ,j2
X X X
= ∆1 − Fρi (1,··· ) ∆ρi (1,··· ) − Fρi Fρij ∆ρi ρij
i i j
X X
− Fρi ( Fρij Fρi,j ∆¯ρi ρi ρi,j + · · ·)
k j k
i j,k
= ··· ,
where loops ρij1 andρij2 are nontouched to each other but both touched with
loop ρi , ρi,j i
k the loop touched with loops ρi or ρj , ∆ρi ρi ρi,j ··· the system deter- j k
minant by removing all the plants in loops ρi , ρij , ρi,j k , · · · from the system. It
is noted that the set of the closed-paths δi (1, · · · ) each of which passes through
g1 consists of the loops ρi (1, · · · ) and the non-loop closed-paths formed by
ρi (1, · · · ) plus ρij , ρi (1, · · · ) plus ρij and ρi,j
k , · · · . Thus, we have the following
lemma.
P
Lemma 3.1.3. ∆ = ∆1 − Fδi (1,··· ) ∆δi (1,··· ) .
i
For example, there exist four closed-paths through plant g1 in Figure 3.5
and they are δ1 (1, 7), δ2 (1, 2, 5, 7), δ3 (1, 2, 3, 4, 5, 7) and δ4 (1, 2, 3, 6). The cor-
responding gains are Fδ1 = g1 g7 , Fδ2 = g1 g2 g5 g7 , Fδ3 = g1 g2 g3 g4 g5 g7 and
Fδ4 = g1 g2 g3 g6 . Obviously, δ1 = ρ1 and δ4 = ρ2 are loops, but δ2 and δ3 are
not. We have ∆1 = 1 − g2 g5 − g3 g4 , ∆δ1 = 1 − g3 g4 − g2 g5 , ∆δ2 = 1 − g3 g4 ,
∆δ3 = 1 and ∆δ4 = 1. It then follows that
3.1 Internal Stability 83
∆ 1 − Fδ 1 ∆ δ 1 − Fδ 2 ∆ δ 2 − Fδ 3 ∆ δ 3 − Fδ 4 ∆ δ 4
= (1 − g3 g4 − g2 g5 ) − g1 g7 (1 − g3 g4 − g2 g5 ) − g1 g2 g5 g7 (1 − g3 g4 )
− g1 g2 g3 g4 g5 g7 (1) − g1 g2 g3 g6 (1)
= 1 − g3 g4 − g2 g5 − g1 g7 − g1 g2 g3 g6 + g1 g7 g3 g4 ,
= g1 g7 g2 g5 (1 − g3 g4 ) + g1 g7 g2 g5 g3 g4 (1)
= Fδ 2 ∆ δ 2 + F δ 3 ∆ δ 3 ,
and
Fρ2 (∆¯ρ2 − ∆ρ2 ) = g1 g2 g3 g6 (1 − 1) = 0,
so that
With Lemmas 3.1.1 and 3.1.3, we are now in the position to prove Theo-
rem 3.1.2.
∆ = det(W ). (3.25)
We will prove (3.25) by induction. If the system has only one plant, then
det(W ) = 1 and ∆ = 1, (3.25) is thus true. When the system contains two
plants g1 and g2 , then n = 2 and there are only two cases to be considered.
Case 1 is that these two plants form a loop, and then its connection matrix
is
1 −g1
W = ,
−g2 1
and det(W ) = 1 − g1 g2 . From (3.8), ∆ = 1 − g1 g2 . (3.25) thus holds in this
case. Case 2 is that these two plants do not form a loop, and then we have
10
W = ,
01
and det(W ) = ∆ = 1. (3.25) is thus proven for the case of two plants. Assume
now that (3.25) holds for systems with (n − 1) plants, we need to show that
it is also true for the system with n plants. The above assumption implies
that det(W1 ) = ∆1 for (n − 1) plants and det(Wδi ) = ∆δi for (n − 1) plants.
Substituting them into (3.22) yields
84 3. Stability and Robustness of Feedback Systems
X
det(W ) = ∆1 − Fδi(1,··· ) ∆δi(1,··· )
i
= ∆,
where the last equality follows from Lemma 3.1.3. Hence (3.25) holds true
for any integer n. 2
In this section, two theorems on internal stability of interconnected sys-
tems, one for scalar signals and other for vector signals, have been established.
They are stated in terms of a single polynomial only. This polynomial is read-
ily found from a given system and can be read off by inspection in many cases.
The application to a few typical control systems has shown simplicity and
effectiveness of the results.
V (s)
+
R (s) + E (s) U (s ) Y (s )
K (s) G (s)
_ +
−1
Let G = D−1 N = NR DR be coprime PMFs of G and K = Y X −1 =
XL−1 YLbe coprime PMFs of K, pG and pK be the characteristic polynomials
of G and K, respectively.
Theorem 3.2.1. The following are equivalent:
(i) The system in Figure 3.10 is internally stable;
(ii) pc := pG pK det[I + GK] = pG pK det[I + KG] is a stable polynomial;
(iii) det[DX + N Y ] is a stable polynomial, or det[XL DR + YL NR ] is a
stable polynomial;
(iv) H(s) is stable.
Proof. (i) ⇐⇒ (ii) It follows from Example 3.1.1 for the case of n = 2.
subtract the last block row from the second block row to get
DX N
−Y 0
X 0 ,
0 I
which is ensured to have full column rank due to the assumed coprimeness
of X and Y . Thus,
−1
X0 DX N
0 I −Y I
Notice
DX N DX N I 0
det = det det
−Y I −Y I Y I
DX N I 0
= det
−Y I Y I
DX + N Y N
= det
0 I
= det[DX + N Y ].
The fact that stable det[DX + N Y ] implies stability of H can also be seen
in an explicit way as follows. By the following well-known matrix identities,
and
(I + KG)−1 = I − (I + GK)−1 KG
= I − K(I + GK)−1 G,
(I + GK)−1 −(I + GK)−1 G
H(s) = , (3.31)
K(I + GK)−1 I − K(I + GK)−1 G
"1 #
s 0
K(s) = 1
.
1 s
−1
10 s0
K(s) = .
s1 os
s3 +4s2 +6s+3
= s(s+1)(s+2) .
88 3. Stability and Robustness of Feedback Systems
We have
pc (s) = s(s3 + 4s2 + 6s + 3),
a unstable polynomial, so the system is unstable. One can easily verify that
det[DX + N Y ] = pc (s). One also find
" 1 1 #
s+1 s(s+1)
GK = 1 2
,
s+2 s(s+2)
H21 = (I + KG)−1 K
3 2
s +3s +4s+3 s+2
s(s3 +4s2 +6s+3) − s(s3 +4s 2 +6s+3)
= ,
s3 +4s2 +7s+4 1
s3 +4s2 +6s+3 s
which is unstable. ♦
Instability of this example is caused by some closed right half plane (RHP)
pole-zero cancellation between G and K. Indeed,
−1 −1
s+1 0 s1 s0
GK =
0 s+2 s2 0s
−1 −1
s+1 0 11 10
=
0 s+2 12 0s
indicates a RHP pole-zero cancellation at s = 0 between G and K.
Another cause for instability is that some RHP pole is lost when forming
det[I + GK], as can be seen in the following example.
Example 3.2.2. Let
" 1 #
s+1 0 10
G(s) = and K= .
1 1 01
s−1 s+1
_
u1 y1
1
1
s +1
1
s- 1
u2 + y2
1
1
_ s +1 +
If K(s) and G(s) are both unstable, then it is necessary to check all four
blocks of the H-matrix to ensure internal stability. However, if at least one
of them is stable, then there is less work to do.
Theorem 3.2.2. If K(s) is stable, then the feedback system shown in Figure
3.10 is internally stable if and only if H12 (s) = −[I + G(s)K(s)]−1 G(s) is
stable.
Proof. “only if”: Since all Hij are stable, H12 is stable, too.
“if”: It follows from (3.29) that
H11 = (I + GK)−1
= (I + GK)−1 (I + GK − GK)
= I + H12 K
is stable since I, H12 and K are so. By (3.31), H21 = K(I + GK)−1 = KH11
is stable, and H22 = I − K(I + GK)−1 G = I − KH12 is stable, too.
where D(s) ∈ Rp×p [s], N (s) ∈ Rp×m [s], and P (s) ∈ Rp×p [s] are given (with
P (s) being stable in contents of control engineering), and X(s) ∈ Rp×p [s]
and Y (s) ∈ Rm×p [s] are sought to satisfy the equation. Such an equation is
called Diophantine equation. Equation (3.32) is linear in X and Y and this
enables us to establish the following results.
90 3. Stability and Robustness of Feedback Systems
Proof. (i) Let L(s) be a GCLD of D(s) and N (s). Then, it follows that there
is a unimodular U (s) such that
[D N ]U = [L 0]
DU1 + N U2 = L (3.34)
for some polynomial matrices U1 and U2 . If L(s) is also a left divisor of P (s),
i.e. P (s) = L(s)P (s) for some polynomial matrix P , then post-multiplying
(3.34) by P yields
showing that X(s) = U1 P and Y (s) = U2 P meets (3.32), and thus form a
solution pair. Conversely, if (3.32) has a solution (X, Y ), it can be written as
LDX + LN Y = P,
or
L(DX + N Y ) = P,
DX0 + N Y0 = P. (3.35)
D(X − X0 ) + N (Y − Y0 ) = 0,
or
3.2 Unity Feedback Systems 91
X − X0
DN = 0. (3.36)
Y − Y0
Since [D N] is of size p × (p + m) and rank p, the right null space of [D N]
is spanned by a (p + m) × m matrix B(s) such that [D N]B = 0 and B has
rank m. The general solution of the homogeneous equation (3.36) can then
be expressed as
X − X0
= BT (3.37)
Y − Y0
bD
for an arbitrary polynomial matrix T . From D−1 N = N b −1 , we have
" #
N b
D N b = 0.
−D
b and D,
And, due to the assumed coprimeness of N b [N
bT − D
b T ]T has full rank
m. Thus, we can take
" #
b
N
B= b ,
−D
so as to reduce its degree as much as possible. Repeat the process for all other
columns and the result would be (Xmin , Ymin ).
With obvious changes, we may work with the case where (X, Y ) is given
and (D, N ) is sought and get the dual results to Theorem 3.2.3.
Theorem 3.2.3 lays a foundation for stabilization and arbitrary pole as-
signment for any plant G(s). One may get its left coprime fraction D−1 (s)N (s)
92 3. Stability and Robustness of Feedback Systems
and specify a suitable P (s) which reflects the designer’s control specifications.
Then, (3.32) is solved to get a solution (X, Y ) and a stabilizer for G(s) is
obtained as K(s) = Y X −1 provided that X is nonsingular and Y X −1 proper.
This procedure actually does more than stabilization or pole assignment be-
cause it can assign an entire polynomial matrix P (s) instead of just det P (s),
the characteristic polynomial of the feedback system.
and
X
∆ arg det(I + kG(s)) = ∆ arg[1 + kλi (s)]. (3.39)
i
The Nyquist plot of λi (s) are called the characteristic loci of G(s). They
have m branches for an m × m plant G(s). It follows from (3.39) that we can
infer closed-loop stability by counting the total number of encirclements of
the point (−1 + j0) made by the characteristic loci kλi (jw) of kG(s), which
is as powerful as in the classical SISO Nyquist criterion as it is now enough
to draw λi (jw) once, usually for k = 1.
3.2 Unity Feedback Systems 95
which are shown in Figure 3.12. The open loop is stable, i.e. Po = 0. For the
closed-loop system to be stable, there should be zero net encirclements of the
critical point (− k1 + j0) made by the loci. Thus, one infers that
• for −∞ < −1 k < 0, there is no encirclement and the closed-loop is stable;
• for 0 < −1
k < 1, there are one or two clockwise encirclements, implying
closed-loop instability; and
• for −1
k > 1, there is no encirclement, and closed-loop stability is ensured.
♦
Im
0.5
0.4
0.3
0.2
0.1
0.5 Re
0
0. 1
0. 2
0. 3
0. 4
0. 5
They also lie in the union of circles, each with centre zii and column radius:
m
X
|zji |, i = 1, 2, . . . , m.
j=1,j6=i
Consider only the i − th row for which |vi | ≥ |vj | for all j 6= i,
m
X
zij vj = λvi ,
j=1
or
m
X
λvi − zii vi = zij vj .
j=1,j6=i
It follows that
m
X m
X m
X
vj vj
|λ − zii | = | zij |≤ |zij | · | |≤ |zij |,
vi vi
j=1,j6=i j=1,j6=i j=1,j6=i
Consider now a proper rational matrix G(s). Its elements are functions
of a complex variable s. The Nyquist array of G(s) is an array of graphs, the
(i, j) graph of which is the Nyquist plot of gij (s). For any i, at each s = jw,
a circle of either row radius:
m
X
|gij (jw)|,
j=1,j6=i
or column radius
m
X
|gji (jw)|,
j=1,j6=i
3.2 Unity Feedback Systems 97
is superimposed on gii (jw). The “bands” obtained in this way are called the
Gershgorin bands. For illustration, Figure 3.13 shows the Nyquist array and
Gershgorin bands of
" s+2 s+1 #
2s+1 2s+1
G(s) = s+1 s+2
. (3.40)
2s+1 2s+1
1.5 1.5
1 1
0.5 0.5
0 0
−0.5 −0.5
−1 −1
−1.5 −1.5
−2 −2
−1 0 1 2 3 −1 0 1 2 3
element (1,1) element (1,2)
1.5 1.5
1 1
0.5 0.5
0 0
−0.5 −0.5
−1 −1
−1.5 −1.5
−2 −2
−1 0 1 2 3 −1 0 1 2 3
element (2,1) element (2,2)
Graphical Test for Dominance can be easily developed. Draw the Gersh-
gorin bands for G(s). If the banks with row radius exclude the origin, then
G(s) is row diagonally dominant. If it is the case with column radius, G(s)
is column diagonally dominant. A greater degree of dominance corresponds
to narrower Gershgorin bands and more closely resembles m decoupled SISO
systems. For G(s) in (3.40), Figure 3.13 indicates that G(s) is diagonally
dominant, which can be confirmed analytically by noting |jw + 2| > |jw + 1|
for any real w.
For diagonally dominant G(s), we have the following stability criterion.
for each i and for all s on the Nyquist contour, that is I + GK is column
diagonally dominant. Let the Nyquist plot of gii (s) encircle the point, (−1/ki +
j0), Ni times anticlockwise. Then, the system is stable if and only if
m
X
Ni = Po .
i=1
Proof. Due to the assumed column diagonal dominance, the following three
numbers are equal:
Pm• the total encirclements of (−1/ki + j0) made by Gii (jw) for all i, N =
i=1 Ni ,
• the total encirclements of (−1/ki + j0) made by Gershgorin bands,
• the total encirclements of (−1/ki + j0) made by characteristic loci.
Hence, the result follows from the generalized Nyquist Stability Criterion.
3.3 Plant Uncertainty and Stability Robustness 99
where δ(·) is a positive scalar function. Note that δ represents the absolute
magnitude of model errors in the additive perturbation case and the relative
magnitude in the multiplicative perturbation cases.
The design problem in an uncertain environment consists in selecting a
controller K(s) in a unity feedback system so as to ensure stability as well as
satisfactory performances not only for the nominal plant G0 but also for any
G which belongs to the given uncertain model set. As for the basic stability
requirement, the control system is called robustly stable if it is stable for all
permissible G.
Proof. We will prove (i) only as other two cases are similar. Since the closed-
loop is stable for ∆(s) = 0 (nominally stabilized), and since G and G0 have
been assumed to have the same number of unstable poles, it follows from The-
orem 3.2.5 that the closed-loop remains stable as long as there is no change
in the number of encirclements of the origin made by det(I + G(jw)K(jw))
with respect to the nominal det(I + G0 (jw)K(jw)), which is equivalent to
det[I + G(jw)K(jw)] 6= 0
Let M = K(I + G0 K)−1 and M (jw) have the singular value decomposition:
M = U ΣV H
The third set of inputs and outputs is novel and comes from uncertainty. For
each uncertainty in the plant (either structured or unstructured), we take it
outside the plant and assign it with one block. Collect them all together as a
special system, which is around the plant and has a block-diagonal structure,
on the diagonal are just those blocks which have been pulled out from inside
the plant. Write
k∆i k∞ ≤ 1, i = 1, 2, · · · , n. (3.52)
or equivalently
In general, however, a uncertainty ∆ may have more than one blocks in its
standard representation in Figure 3.15, i.e. n ≥ 2 for ∆ in (3.52). Then the
condition (3.54) is only sufficient for robust stability, because most perturba-
tions which satisfy k∆k ≤ 1 are no longer permissible. (3.54) is usually too
conservative to use for such structured uncertainties.
Let BDδ denote the set of stable, block-diagonal perturbations with a
particular structure, and k∆j k ≤ δ. Suppose nominal stability of the system
in Figure 3.15. It follows from the multivariable Nyquist theorem that the
feedback system can be unstable if and only if
for some w and some ∆ ∈ BDδ . We thus define the structured singular value
µ(Q22 (jw)) as follows:
µ(Q22 (jw)) =
0, if det[I − Q22 ∆] 6= 0, for any ∆ ∈ BD∞ ;
(3.56)
1/(min∆∈BD∞ {σ(∆(jw))| det[I − Q22 (jw)∆(jw)] = 0}), otherwise.
In other words, one tries to find the smallest structured ∆ (in the sense of
1
σ(∆)) which makes det(I − Q22 (jw)) = 0, then, µ(Q22 ) = σ(∆) ; If no ∆
meets (3.55), then µ(Q22 ) = 0.
Example 3.3.2. Let
2 2
Q22 = .
−1 −1
The smallest (unstructured) full matrix perturbation ∆ which makes det(I −
Q22 ∆) = 0 is
0.2 0.2
∆= ,
−0.1 −0.1
and
This example shows that the structured singular value µ(.) depends on
the structure of ∆ and that a larger perturbation in the structured form is
usually allowed than unstructured one, before the matrix becomes singular.
One notes that if ∆ = 0, I − Q22 ∆ = I is nonsingular. Thus, to obtain
µ(Q22 ), one way is to gradually increase σ(∆) until we find I − Q22 ∆ is
singular. Besides, the sequence of Q22 and ∆ in µ-definition does not matter
since
The structured singular value µ(Q) depends on the structure of the set BDδ
as well as on Q. It has the following properties.
One sees that this condition is similar to (3.54) for the unstructured case.
The only difference is that the singular value has been replaced here with
the structured singular value. Under condition (3.57), one can show that
maxω |λmax (Q22 ∆)| < 1 so that (I − Q22 ∆) can not be singular on ω axis and
det(I − Q22 (jw)∆(jw)) 6= 0, and the system is thus stable for all ∆ meeting
k∆j k ≤ 1. If ∆ has one block only, by Lemma 3.3.3(ii), (3.57) reduces to
(3.54). It should be pointed out that the structured singular value is difficult
to compute and we usually use the property (iv) in Lemma 3.6 to overestimate
µ(Q).
Example 3.3.3. Continue Example 3.3.1, we have to find Q22 which meets
x = Q22 z. One sees
108 3. Stability and Robustness of Feedback Systems
x1 α11 u1 α11 0
x2 α21 u1 α21 0
x= = = u1 ,
x3 α12 u2 0 α12 u2
x4 α22 u2 0 α22
z1
u1 y1 u1 1010 z2
= −K = −K G0 + ,
u2 y2
u2 0 1 0 1 z3
z4
giving
u1 −1 1010
= −(I + KG0 ) K z,
u2 0101
so that
α11 0
α21 0
x = − (I + KG0 )−1 K 1 0 1 0 z.
0 α12 0101
0 α22
:= Q22 z.
The system remains stable if and only if
kQ22 kµ ≤ 1,
Theorems 3.3.1 and 3.3.2 are actually special cases of the small gain theo-
rem. Essentially, they all say the same thing: if a feedback loop consists of
stable systems and if the open-loop gain (in a suitable sense) is less than
unity, then the closed-loop system is stable. These results are elegant and
have been widely used in the area of robust control. Unfortunately, it could
be very conservative in many cases. To reveal this conservativeness, one notes
that descriptions in (3.46) for unstructured uncertainty and (3.51) and (3.52)
for structured ones are actually the same for each individual block ∆i which
is restricted by its bound only, and allow much richer uncertainties to occur
than what could exist in reality. In particular, phase changes in the uncer-
tainties are not limited at all, and an infinite phase change is permissible,
which is definitely unrealistic in practice. In real life, both gain and phase
uncertainties would be finite. For example, a common practice in modelling is
that a parametric plant model with some parameter uncertainty is in general
valid over a working frequency range, and may become useless beyond the
3.3 Plant Uncertainty and Stability Robustness 109
range due to unmoulded dynamics, noise and so on. This implies that phase
uncertainty is much smaller in low frequencies than that in high frequencies
range, and surely finite. For illustration, let us give a concrete example to
show conservativeness of Theorem 3.3.1 and to motivate our sequel develop-
ment.
where δ1 ∈ [−0.5, 0.5], δ2 ∈ [−0.5, 0.5] and δ3 ∈ [−0.5, 0.5]. Suppose that
the controller K(s) = 1. One may easily see that the closed-loop system is
robustly stable, because its characteristic polynomial,
p(s) = s2 + (1 + δ2 )s + 2 + δ1 + δ3 , (3.59)
always has positive coefficients and thus its roots lie in th open left half of
the complex plane.
Now, Let us try the small gain theorem. The nominal plant is
1
G0 (s) = . (3.60)
s2 + s + 1
we may express the uncertainty as an additive one:
1 + δ1 1
∆(s) = − . (3.61)
s2 + (1 + δ2 )s + (1 + δ3 ) s2 + s + 1
Theorem 3.3.1 would say that the feedback system is robustly stable if and
only if
where
−K(s) s2 + s + 1
Q22 (s) := =− 2 . (3.63)
1 + K(s)G0 (s) s +s+2
We have
1 1 + δ1
sup|Q22 (jω)∆(jω)| > lim |Q22 (jω)∆(jω)| = ×| − 1|. (3.64)
ω ω→0 2 1 + δ3
If we choose δ1 = 0.5 and δ3 = −0.5, or the plant is perturbed to
1.5
G(s) = , (3.65)
s2 + s + 0.5
there results in
110 3. Stability and Robustness of Feedback Systems
violating (3.62) though the system is stable. This shows the conservativeness
of (3.62) that assumes infinite phase change at any frequency, which is not
true in this example. ♦
It should be pointed out that the structured uncertainty and the theory of
structured singular values cater to the block diagonal structure of the uncer-
tainty, and have nothing to do with further information on individual blocks
such as breaking them into magnitude and phase. In the rest of this subsec-
tion, we will acknowledge the importance of limited phase uncertainty and
incorporate such information into the robust stability conditions, in addition
to the existing magnitude information, so as to remove conservativeness and
make the stability criterion exact.
To be specific, we take the SISO unity feedback system for study. If the
plant G(s) and the controller K(s) have no poles in the closed right half
plane, then the Nyquist criterion tells that the closed-loop system is stable
if and only if the Nyquist curve of G(jω)K(jω) does not encircle the critical
point, (−1 + j0). An uncertain plant is always represented by a set of models.
Each model gives rise to a Nyquist curve and the set thus sweeps a band of
Nyquist curves. Then from the Nyquist criterion, the closed-loop system is
robustly stable if and only if the band of the Nyquist cluster for the uncertain
plant and the controller does not encircle the point, (−1 + j0). Conversely,
the system is not robustly stable if and only if there exits a Nyquist curve
among the cluster which encircles −1 point, or there must exist a specific
frequency ω0 at which the frequency response region of the uncertain loop is
such that both
and
and
Theorem 3.3.3. Suppose stability of the uncertain plant and the controller.
The uncertain closed-loop system is robustly stable if for any ω ∈ R, either
1
|K(jω)| 6 (3.71)
max{|G(jω)|}
G
or
holds.
Corollary 3.3.1. For the system in Figure 3.15, if the nominal controlled
system Q22 (s) and the uncertainty ∆(s) are both stable, then the uncertain
closed-loop system remains stable if for any ω ∈ R, either
1
|Q22 (jω)| 6 (3.73)
max{|∆(jω)|}
∆
or
holds.
For analysis of a given uncertain control system, the robust stability cri-
terion in (3.71) and (3.72) can be tested graphically as follows:
(i) Draw Bode plots for magnitude 1/max{|G(jω)|} and phase min{arg
{G(jω)}} respectively, according to the given structure of the uncertain
plant;
(ii) Draw Bode plots for K(jω) in the same diagram; and
(iii) Check if condition (3.71) or (3.72) is satisfied.
Similarly, one can graphically determine the boundaries of the Nyquest band
of a given uncertain plant. Such boundaries will contain the worst case of the
plant which can facilitate robust stability analysis and robust control design.
Obviously, it is more powerful if analytical expressions for the uncertain plant
boundaries can be derived. Though this is difficult in general, it is possible
to do so for a class of typical processes, the uncertain second-order plus dead
time model.
112 3. Stability and Robustness of Feedback Systems
Let us consider
b
G(s) = e−Ls , (3.75)
s2 + a1 s + a2
and
q
b− a−
2 +a2
+
√ , ω ∈ 0, 2 ;
(a1 ω) +(a+
+ 2 2 2
2 −ω )
min{|G(jω)|} = q (3.78)
b− a−
2 +a2
+
√ , ω∈ 2 ,∞ .
+
(a1 ω) +(ω 2 −a−
2
2 )
2
It follows that
q
−arg(a− 2 + +
2 − ω + ja1 ω) − L ω,ω ∈ 0, a− 2 ;
min{arg(G(jω))} = q
−arg(ja−
1 ω + a−
2 − ω 2
) − L +
ω, ω ∈ a−
2 , ∞ ;
and
q
+ −
−arg(a 2 − ω 2
+ ja1 ω) − L −
ω, ω ∈ 0, a+ 2 ;
max{arg(G(jω))} = q
+ + 2 − +
−arg(ja 1 ω + a2 − ω ) − L ω, ω ∈ a2 , ∞ .
(3.80)
3.4 Notes and References 113
0
−1 0 1
10 10 10
−50
−100
−150
−200
−1 0 1
10 10 10
Example 3.3.5. In Example 3.3.4, the small gain theory fails to decide sta-
bility of the uncertain system given there. We now use Theorem 3.3.3, an
exact robust stability criterion, with the help of the above precise uncertainty
bounds to re-examine the system. Apply (3.77) and (3.80) to get
√
√ 1.5
, ω ∈ 0, 0.5 ;
(0.5ω) +(0.5−ω )
2 2 2
√ √
1.5
max{|G(jω)|} = 0.5ω , ω∈ 0.5, 1.5 ;
√
√ 1.5
, ω∈ 1.5, ∞ ;
2 2 2
(0.5ω) +(1.5−ω )
and
( √
−arg(0.5 − ω 2 + j1.5ω), ω ∈ 0, 0.5 ;
min{arg(G(jω))} = √
−arg(0.5 − ω 2 + j0.5ω), ω ∈ 0.5, ∞ .
with K(jω) = 1, though the gain condition (3.71) is violated when ω < 1.66,
see the gain plot of Figure 3.17. But the phase condition (3.72) will be satisfied
for all frequencies, see the phase plot of Figure 3.17. Thus the uncertain
system is robustly stable. ♦
In this chapter we shall consider a system of the following state space form
ẋ = Ax + Bu,
y = Cx. (4.1)
If x(0) = 0 and the input and the output vectors have the same dimension
m, these are related by the transfer function matrix:
y(s) = G(s)u(s) = C(sI − A)−1 Bu(s), (4.2)
which may be expanded into
y1 (s) = g11 (s)u1 (s) + g12 (s)u2 (s) + · · · + g1m (s)um (s),
y2 (s) = g21 (s)u1 (s) + g22 (s)u2 (s) + · · · + g2m (s)um (s),
.. (4.3)
.
ym (s) = gm1 (s)u1 (s) + gm2 (s)u2 (s) + · · · + gmm (s)um (s).
These equations are said to be coupled , since each individual input influences
all of the outputs. If it is necessary to adjust one of the outputs without
affecting any of the others, determining appropriate inputs u1 , u2 , . . . , um
will be a difficult task in general. Consequently, there is considerable interest
in designing control laws which remove this coupling, so that each input
control only the corresponding output. A system of the form (4.1) is said to
be (dynamically) decoupled (or non-interacting) if its transfer function matrix
G(s) is diagonal and non-singular, that is,
y1 (s) = g11 (s)u1 (s),
y2 (s) = g22 (s)u2 (s),
.. (4.4)
.
ym (s) = gmm (s)um (s).
and none of the hii (s) are identically zero. Such a system may be viewed as
consisting of m independent subsystems. Note that this definition depends
upon the ordering of inputs and outputs, which is of course quite arbitrary.
The definition will be extended to block-decoupling case in Section 4.5.
One example of a system requiring decoupling is a vertical take-off air-
craft. The outputs of interest are pitch angle, horizontal position, and al-
titude, and the control variables consist of three different fan inputs. Since
Q.-G. Wang: Decoupling Control, LNCIS 285, pp. 115-128, 2003.
Springer-Verlag Berlin Heidelberg 2003
116 4. State Space Approach
these are all coupled, the pilot must acquire considerable skill in order to
simultaneously manipulate the three inputs and successfully control the air-
craft. The system may be decoupled using state-variable feedback to provide
the pilot with three independent and highly stable subsystems governing his
pitch angle, horizontal position, and altitude.
It should be pointed out that dynamic decoupling is very demanding. A
signal applied to input ui must control output yi and have no whatsoever
effect on the other outputs. In many cases this requires a complex and highly
sensitive control law, and in other cases it cannot be achieved at all (without
additional compensation). It is useful, therefore, to consider a less stringent
definition which involves only the steady-state behavior of the system.
A system of the form (4.2) is said to be statically decoupled if it is stable
and its static gain matrix G(0) is diagonal and non-singular. This means
that for a step function input u(t) = α1(t), where α = [α1 , α2 , · · · , αm ]T is
constant, the outputs satisfy
Proof. If (A, B) is stabilizable, one can find a K such that (A−BK) is stable.
Let us assume that such a K has been found. Consider
AB AB In 0
rank = rank
CD CD −K Im
A − BK B
= rank
C − DK D
In 0 A − BK B
= rank
(C − DK)(−A + BK)−1 Im C − DK D
A − BK B
= rank . (4.10)
0 (C − DK)(−A + BK)−1 B + D
Design for Static Decoupling. Assume that conditions (i) and (ii) in
Theorem 4.1.1 are satisfied, thus the problem is solvable. We proceed as
follows.
(i) Design a K such that the state feedback system is stable, i.e, det(sI −
A + BK) has all its roots in the left half plane;
(ii) Obtain F as F = [(C − DK)(−A + BK)−1 B + D]−1 ; and
(iii) Form the feedback control law as u = −Kx + F r.
118 4. State Space Approach
Therefore decoupling by state feedback requires one to find the control ma-
trices K and F which make H(s) diagonal and non-singular. Before treating
the decoupling problem, we first consider the relation between the transfer
function H(s) and that of the plant (4.11).
G(s) = C(sI − A)−1 B. (4.16)
Proposition 4.2.1 indicates that controlling the system (4.11) with the
control law (4.12) is equivalent to compensating for the system (4.11) serially
by using the compensator
Hc (s) = [I − K(sI − A + BK)−1 B]F, (4.19)
as shown in Figure 4.1. This compensator can be represented by the state
space equations:
ẋc = (A − BK)xc + BF r, (4.20)
yc = −Kxc + F r, (4.21)
where r is the input to the compensator and yc is its output.
120 4. State Space Approach
Let
cT1
cT2
C= . ,
..
cTm
and define integers σi , i = 1, 2, · · · , m, by
ÿ
min j|cTi Aj−1 B 6= 0T , j = 1, 2, · · · , n − 1 ;
σi = (4.22)
n − 1; if cTi Aj−1 B = 0T , j = 1, 2, · · · , n.
To actually evaluate σi , we start with i = 1, cTi = cT1 . Let j = 1, then
cTi Aj−1 B = cT1 A0 B = cT1 B. If cT1 B = 0, then σ1 = 1; otherwise, increase j
to 2, if cT1 Aj−1 B = c1 AB = 0, then σ1 = 2; otherwise, increase j again until
j = n. Repeat this procedure for i = 2, 3, · · · , m.
With σi , i = 1, 2, · · · , m, we can define
T σ −1
c1 A 1 B
cT2 Aσ2 −1 B
B∗ = .. , (4.23)
.
cTm Aσm −1 B
and
c1 Aσ1
c2 Aσ2
C∗ = .. . (4.24)
.
cm Aσm
Theorem 4.2.1. There exists a control law of the form (4.12) to decouple
the system (4.11) if and only if the matrix B ∗ is non-singular. If this is the
case, by choosing
4.2 Dynamic Decoupling 121
F = (B ∗ )−1 ,
K = (B ∗ )−1 C ∗ ,
the resultant feedback system has the transfer function matrix:
H(s) = diag{s−σ1 , s−σ2 , · · · , s−σm }.
Proof. Necessity is proved first. Noting that
(sI − A)−1 = s−1 I + As−2 + A2 s−3 + · · · , kA/sk < 1,
the i-th row of the transfer function G(s) of (4.16) can be expanded in poly-
nomials of s−1 as
gi (s)T = cTi (sI − A)−1 B = cTi Bs−1 + cTi ABs−2 + · · ·
= cTi Aσi −1 Bs−σi + cTi Aσi Bs−(σi +1) + · · ·
= s−σi (cTi Aσi −1 B + cTi Aσi Bs−1 + cTi Aσi +1 Bs−2 + · · · )
= s−σi {cTi Aσi −1 B + cTi Aσi (sI − A)−1 B}. (4.25)
Using B ∗ of (4.23), G(s) can be written as
−σ1
s 0 ··· 0
.
0 s−σ2 . . . .. ∗
G(s) = .
[B + C ∗ (sI − A)−1 B].
(4.26)
.. . .
.. .. 0
0 · · · 0 s−σm
Thus from Proposition 4.2.1 , the transfer function matrix of the feedback
system is given by
H(s) = G(s)[I − K(sI − A + BK)−1 B]F
= diag(s−σ1 , s−σ2 , . . . , s−σm )[B ∗ + C ∗ (sI − A)−1 B]
× [I − K(sI − A + BK)−1 B]F. (4.27)
∗ ∗ −1
For decoupling, H(s) is non-singular, so are [B + C (sI − A) B][I −
K(sI − A + BK)−1 B]F , and its coefficient matrix for s0 , B ∗ F . Thus the
non-singularity of B ∗ is a necessary condition for decoupling.
The sufficiency is proved by constructing the control law to decouple the
system as follows. If B ∗ is non-singular, let K and F be chosen as
K = B ∗−1 C ∗ , (4.28)
∗−1
F =B . (4.29)
It follows from Proposition 4.2.1 that
H(s) = diag (s−σ1 , s−σ2 , . . . , s−σm )[B ∗ + C ∗ (sI − A)−1 B]
×[F −1 + F −1 K(sI − A)−1 B]−1 . (4.30)
Substituting (4.29) and (4.28) into (4.30) gives
H(s) = diag(s−σ1 , s−σ2 , . . . , s−σm ), (4.31)
and the system is decoupled.
122 4. State Space Approach
The form of the decoupled system (4.31) having s−σi as the diagonal
elements is called an integrator decoupled system.
yielding σ1 = 1; And
11
cT2 A0 B = cT2 B = 0 1 = 0,
00
00 11
cT2 A1 B = cT2 AB = 0 1 = 11 ,
10 00
giving σ2 = 2. But one has
T σ −1 T
c A 1 B c B 11
B ∗ = 1T σ2 −1 = T1 = ,
c2 A B c2 AB 11
which is singular, and the system cannot be decoupled.
It is noted that the B-matrix of the above system is of rank defect. It is
trivial to show from the feedback system transfer function matrix H(s) in
(4.15) that a necessary condition for decoupling is that both C and B have
full rank. ♦
so that
10
B∗ = .
43
4.2 Dynamic Decoupling 123
Note that
cT AB
∗ 10
B = 1T = =I
c2 AB 01
is non-singular. Thus decoupling can be achieved with
2
∗−1 ∗ ∗−1 3ω 0 0 2ω
u(t) = −B C x(t) + B v(t) = − x(t) + v(t),
0 −2ω 0 0
which results in the closed-loop system:
0 1 0 0 0 0
0 0 0 0 1 0
ẋ = (A − BB ∗−1 C ∗ )x + Bv = 0
x + v,
0 0 1 0 0
0 0 0 0 0 1
1000
y = Cx = x.
0010
Note that this decoupled system is in the controllable canonical form, and
we may apply additional feedback:
k1 k2 0 0 v̄
v = Kx + v̄ = x+ r ,
0 0 k3 k4 v̄θ
to have
0 1 0 0 0 0
k1 k2 0 0 1 0
ẋ = (A − BB C + BK)x + v̄ =
∗−1 ∗
0
x + v̄,
0 0 1 0 0
0 0 k3 k4 0 1
1000
y = Cx = x.
0010
If we choose the feedback gains k1 = k3 = −1 and k2 = k4 = −2, each of
the two independent subsystems will be stable with a double pole at s = −1.
The transfer function matrix may easily be shown to be
1 0
2
(s + 1)
H(s) = .
0 1
2
(s + 1)
This closed-loop system is not only stable (perturbations from the nominal
trajectory will always decay to zero), but adjustments to r and θ can be made
independently. ♦
In this section, the control gain matrices K and F are determined so that
the decoupled system has pre-assigned poles. To this end, let
m
X
n̄ = σi
i=1
Theorem 4.3.1. Suppose that the system (4.11) can be decoupled by state
feedback. If K and F are chosen as
F = (B ∗ )−1 ,
(4.32)
K = (B ∗ )−1 C ∗∗ ,
then the resultant feedback system has the transfer function given by
1 1 1
H(s) = diag , ,··· , . (4.33)
φ1 (s) φ2 (s) φm (s)
Proof. Let the i-th row vector of the transfer function matrix of (4.16) be
denoted by giT (s). Multiplying it by (sσi + · · · + γiσi ) yields from (4.25)
126 4. State Space Approach
" #
1 0
H(s) = s .
0 1s
However, if we want to decouple the system to get
H(s) = diag[(s + 1)−1 , (s + 2)−1 ],
then, we have γ11 = 1, γ21 = 2. (4.32) becomes
T
c1 A γ11 cT1
K= T +
c2 A γ21 cT2
T
1c1
= C∗ +
2cT2
0 0 1 110 1 1 1
= + = ,
−1 −2 −3 002 −1 −2 −1
and F = I remains the same as before. ♦
Theorem 4.3.2. The linear system (4.11) can be decoupled by the feedback
control law (4.12) with the i-th diagonal element of the transfer function of
the decoupled system having a numerator ni (s) if and only if the i-th row of
G(s), denoted by giT (s), is represented by
and
c̄T1 A(σ̄1 −1) B
T (σ̄ −1)
c̄2 A 2 B
B̄ =
∗
..
(4.38)
.
c̄Tm A(σ̄m −1) B
On the other hand, it follows from (4.17), (4.41) and (4.42) that
cTi (sT − A)−1 B[I + K(sI − A)−1 B]−1 F
= ni (s)c̄Ti (sI − A)−1 B[I + K(sI − A)−1 B]F.
(4.39) is then derived by dividing both sides of the above by [I + K(sI −
A)−1 B]−1 F from the right side.
The above theorem indicates that the numerator terms in the decoupled
system are given by the properties of the original system and cannot be
changed.
In contrast to the preceding chapter where a state space model of the plant
and state feedback controller are exclusively utilized for decoupling problem,
we will employ unity output feedback compensation to decouple the plant in
this chapter. Polynomial matrix fractions as an input-output representation
of the plant appear a natural and effective approach to dealing with output
feedback decoupling problem. The resulting output feedback controller can
be directly implemented without any need to construct a state estimator.
In particular, we will formulate a general decoupling problem and give some
preliminary results in Section 1. We start our journey with the diagonal de-
coupling problem for square plants in Section 2 and then extend the results
to the general block decoupling case for possible non-square plants in Sec-
tion 3. A unified and independent solution is also presented in Section 4. In
all the cases, stability is included in the discussion. A necessary and suffi-
cient condition for solvability of the given problem is first given, and the set
of all compensators solving the problem is then characterized. Performance
limitations of decoupled systems are compared with a synthesis without de-
coupling. Numerous examples are presented to illustrate the relevant concepts
and results.
T = GK[I + GK]−1
T11 0
T22
= .
..
0 Tvv
= block diag{T11 , T22 , · · · , Tvv }, (5.1)
with Tii ∈ Rm p
i ×mi
(s) being nonsingular for all i, i = 1, 2, · · · , v. The case
of v = m and m1 = m2 = · · · = mv = 1 is called diagonal decoupling. If
the list of positive integers {mi } is obviously known from the context and no
confusion arises, we denote by Rm×m d [s], Rm×m
d (s), Rm×m
dp (s) and Rm×m dsp (s)
m×m m×m
the set of block diagonal and nonsingular elements in R [s], R (s),
Rm×m
p (s) and Rm×msp (s), respectively. The noninteracting control problem to
be solved in this chapter is stated as follows.
Lemma 5.1.1. Any full row rank matrix G ∈ Rm×l (s) has a stability fac-
torization G = [G+ 0]G− . The matrix G+ is unique up to a multiplication
by an unimodular polynomial matrix from the right. In the case of square
G, G− is unique up to a multiplication by an unimodular polynomial matrix
from the left.
G = U1 SU2 ,
G+ = U1 diag{ri+ },
and
G = [G+ − + −
1 0]G1 = [G2 0]G2 .
We have
[G+ + − − −1
1 0] = [G2 0]G2 (G1 ) .
Define U = G− − −1
2 (G1 ) and partition it as
U11 U12
U=
U21 U22
[G+ + +
1 0] = [G2 U11 G2 U12 ]
and
G+ +
1 = G2 U11 . (5.2)
Note that all the distinct poles of U11 must be the poles of U = G− − −1
2 (G1 )
which is stable. So U11 can only have stable poles if any, and by (5.2) these
poles must be also the poles of G+1 because they cannot be cancelled by G2 .
+
+
But G1 cannot have any stable poles by definition, and thus U11 must not
have any pole. By a similar argument using G+ + −1 −1
2 = G1 U11 , U11 must not
132 5. Polynomial Matrix Approach
have any pole either. Hence, U11 is unimodular. For square G, [G+ 0] reduces
to G+ only and U11 = U . U G− −
1 = G2 and the proof is completed.
Note that for nonsquare full row rank G, G− is highly non-unique: its last
l − m rows are arbitrary subject to the constraint that no poles and zeros
are introduced. The following special stability factorization will be required
later.
Corollary 5.1.1. There exist unique square polynomial matrices Ng+ and
Dg+ and a matrix G− such that G+ + + −1
g := Ng (Dg ) and G− define a stability
factorization, Ng is in the Hermite column form and Dg+ is in the Hermite
+
row form.
Proof. Let G = [G+ − +
1 0]G1 be a stability factorization and G1 = N (D )
+ + −1
T −1 = I + (GK)−1 .
Theorem 5.1.1. For a plant G ∈ Rm×l p (s), (P) is solvable if and only if
there is a K ∈ Rl×m (s) such that GK is block-diagonal and nonsingular, and
has no RHP pole-zero cancellations.
Definition 5.1.2. Let a full row rank N ∈ Rm×l [s] and a nonsingular D ∈
Rm×m [s] be given. If there are N̄ ∈ Rm×l [s] and D̄ ∈ Rl×l [s] such that
DN = N̄ D̄,
where D and N̄ are left coprime, and N and D̄ right coprime, then N and D
are called as externally skew prime and N̄ and D̄ as internally skew prime.
Let us consider first the simplest case of (P): diagonal decoupling for square
plants, which is abbreviated as (PS) for ease of frequent future reference.
Then G should be square and nonsingular. Our task, by Theorem 5.1.1, is to
find a nonsingular K such that GK is nonsingular and diagonal and has no
RHP pole-zero cancellations.
134 5. Polynomial Matrix Approach
5.2.1 Solution
+
Theorem 5.2.2. (PS) is solvable if and only if Nra and D+ are externally
+ + + +
skew prime and N Nra and D̄la D̄ are coprime.
K = G−1 T, (5.6)
where T ∈ Rm×m
d (s) is given by the coprime polynomial matrix fraction
+ + −1
T = N + Nra
+
(D̄la D̄ ) . (5.7)
and
+
δ + (K) ≤ deg(det(D̄la )). (5.11)
(N2+ )−1 NT 2 with N2+ = (N1+ )−1 N + and NT 2 = (N1+ )−1 NT is a coprime
polynomial matrix fraction and (N2+ )−1 is not cancelled in G−1 T . Therefore,
there must be such a representation
with DK Ñ2+ and NK left coprime. But then this unstable Ñ2+ will disappear
in GK because
−1
GK = [(Ñ2+ )−1 Q(N1+ )−1 ]−1 [(DK Ñ2+ )−1 NK ] = N1+ Q−1 DK NK .
(N + )−1 NT = P
or equivalently
NT = N + P
NT = N + Nra
+
P1 , (5.15)
where Nra+
is a right strict adjoint of N + and P1 ∈ Rm×m
d [s]. Substituting
(5.15) into (5.14) gives
+ −1
K = (G− )−1 D+ Nra D T P1 .
+
The coprimeness of NT and DT implies that Nra and DT are right coprime.
+ −1
Construct a dual left coprime polynomial matrix fraction from Nra DT as
D̃T−1 Ñra
+ + −1
= Nra DT (5.16)
with
+ +
det(Ñra ) = a1 det(Nra ), a1 ∈ R. (5.17)
K becomes
where
+ −1 + −1
P̄2 (N̄ra ) = (Ñra ) P̃2 (5.20)
and
+
N̄ra and D+ are left coprime. (5.22)
+
Equations (5.21) and (5.22) hold due to the left coprimeness of D̃T and Ñra .
+
Because DT is a polynomial matrix and P̄2 and N̄ra are right coprime, it
follows from (5.19) that
+ −1 + +
(N̄ra ) D Nra = D̄+ (5.23)
for some polynomial matrix D̄+ . Equation (5.23), together with (5.17), (5.21),
+
and (5.22), implies that Nra and D+ are externally skew prime and N̄ra +
and
+
D̄ just constitute their dual internally skew prime pair. With (5.23), (5.19)
is simplified into
DT = P̄2 D̄+ .
N + Nra
+
P = det(N + )I (5.25)
GK = K1 K2 ,
K = G−1 NT DT−1 ,
+ +
where NT = N + Nra +
P1 , DT = P2 D̄la D̄ for some P1 , P2 ∈ Rm×m
d [s], and
NT and DT are coprime. Define K2 = P1 P2−1 , then NT DT−1 = K1 K2 . The
coprimeness of NT and DT implies that K1 K2 has no RHP pole-zero cancel-
lations. Simple calculations yield
+ −1
K = G−1 K1 K2 = (G− )−1 N̄ra
+
(D̄la ) K2 ,
where K2 ∈ Rm×m d (s) is such that K1 K2 with K1 = N Nra (D̄la D̄)−1 has no
pole-zero cancellations.
(iii) All the achievable open-loop maps under decoupling with arbitrary
pole assignability can be expressed by
GK = K1 K2 ,
G = G+ G− ,
where
1 1
G+ = 1 s ,
s−1 s−1
140 5. Polynomial Matrix Approach
and
1
0
G− = s+1
1 .
0 s+2
One sees that G has a common zero and pole at s = 1, which is unstable.
Compute (N + )−1 as
1
+ −1 s−1 0
(N ) = −1 ,
s−1 1
then,
+ + −1 1 0
Nra = (N ) diag{(s − 1), 1} = .
−1 1
+
Obviously, Nra and D+ are externally skew prime since det(Nra
+
) ∈ R has no
root at all. One easily checks that
s −1
D̄+ =
−1 1
+
and N̄ra = I2 satisfy the skew primeness equation
D+ Nra
+ + +
= N̄ra D̄ ,
and they constitute a dual skew prime pair. A left strict adjoint of D̄+ is
similarly obtained as
+ 1 1
D̄la = .
1 s
Hence, we have
s−1 0
N + Nra
+
= ,
0 1
and
+ + s−1 0
D̄la D̄ = .
0 s−1
They are not coprime. (PS) for this plant is thus not solvable. ♦
5.2 Diagonally Decoupling for Square Plants 141
G is solvable.
Now turn back to the example, we have
G+ = block diag{G+ +
1 , G2 },
where G+
2 = 1/(s − 1) is scalar, for which (PS) is of course solvable, and
s 1
G+1 = 1 1 .
s s
The zero of G+
1 at s = 1 is different from its pole at s = 0, and by Corollary
5.2.1, (PS) for G+1 is also solvable. It follows from Theorem 5.2.4 that (PS)
for G is thus solvable. ♦
search in (i) and thus it is itself a member of Gd,min , say, Gjd,min . Therefore,
we conclude that G̃d,min = Gjd,min D3γ . The proof is completed.
Remark 5.2.1. Note that the locations of poles of a multivariable system must
be the same as those of its elements (Chapter 2), and that the zeros of G
are the poles of G−1 . Thus, we can find the common RHP zeros and poles
of G by examining the poles of the elements of G and G−1 . This observation
together with the proof of Lemma 5.2.2 naturally gives rise to the following
constructive procedure for finding Gd,min from G.
Step 1: Calculate Gd = G−1 and determine the set Γ from pole locations
of elements of G and G−1 . If Γ = φ, i.e. empty set, take Gd,min =
Gd ; otherwise, proceed to step 2.
Step 2: For each γ ∈ Γ , find lpγ and lzγ from pole multiplicity of elements
of Gd and G−1 l1 l2
d . Let Dγ = diag{(s−γ) , (s−γ) , · · · , (s−γ) },
lm
for which we can easily find that the only common RHP pole and zero of G
is at 1 i.e., Γ = {1}. Take Gd = G−1 and obviously we have lpγ = 1 and
lzγ = 1. For D = diag{(s − 1)l1 , (s − 1)l2 }, where −1 ≤ l1 , l2 ≤ 1, the search
for Gd,min is tabulated in Table 5.1, from which two Gd,min are found as
" s(s+1) s+1
#
s−1 − s−1
Gd,min1 = ,
s+2 s+2
− s−1 s−1
and
s(s + 1) −(s + 1)
Gd,min2 = .
−(s + 2) (s + 2)
" s+1 #
1 − s+2
G̃d = s+2
,
− s(s+1) 1
and
" 1 s+1 #
s−1 − (s+2)(s−1)
G̃d,min2 = s+2 1
. (5.30)
− s(s+1)(s−1) s−1
Tables 5.1 and 5.2 verify that δ γ (Gd,min ) is unique for a given G. Although
two sets of the Gd,min obtained above look different, actually they are related
to each other by
−1
s(s + 1) 0
Gd,min1 = G̃d,min2 ,
0 (s + 2)
and
" 1
#−1
s(s+1) 0
Gd,min2 = G̃d,min1 1
,
0 s+2
" s(s+1) #
(s−1)2
−(s + 1)
(−1, 0) s+2 3
− (s−1) 2 s+2
" s(s+1) #
(s−1)2
−(s + 1)(s − 1)
(−1, 1) s+2 4
− (s−1) 2 (s − 1)(s + 2)
" #
s(s+1) s+1
s−1 − s−1
(0, −1) s+2 s+2 1
− s−1 s−1
" #
s(s+1)
s−1 −(s + 1)
(0, 0) s+2 2
− s−1 s+2
" #
s(s+1)
s−1 −(s + 1)(s − 1)
(0, 1) s+2 3
− s−1 (s − 1)(s + 2)
" #
s+1
s(s + 1) − s−1
(1, −1) s+2 2
−(s + 2) s−1
s(s + 1) −(s + 1)
(1, 0) 1
−(s + 2) s + 2
s(s + 1) −(s + 1)(s − 1)
(1, 1) 2
−(s + 2) (s + 2)(s − 1)
sufficiently large integer l such that Gd no longer has any zero (resp. pole) at
ρ. This Gd has no pole-zero cancellation with G at ρ. Repeat this elimination
of pole-zero cancellation for all possible ρ and the resultant Gd will have no
pole-zero cancellation with G at ρ ∈ C+ but ρ ∈ / Γ.
Necessity: It follows from Theorem 5.1.1 that if (PS) is solvable, then
there is a Gd such that GGd is decoupled and meets
" s+1
#
1
s−1 − (s+2)(s−1)
(−1, −1) s+2 1 1
− s(s+1)(s−1) s−1
" #
1 s+1
s−1 − s+2
(−1, 0) s+2 2
− s(s+1)(s−1) 1
" s+1
#
1 − (s−1)(s+2)
(0, −1) s+2 1 2
− s(s+1) s−1
" #
1 − s+1
s+2
(0, 0) s+2 1
− s(s+1) 1
By Lemma 5.2.3, for Gd there exists a Gd,min such that Gd = Gd,min D and
δγ (Gd ) = δγ (Gd,min ) + δγ (D). (5.32)
Substituting (5.32) into (5.31) yields
δγ (G) + δγ (Gd,min ) + δγ (D) = δγ (GGd,min D), (5.33)
implying that G and Gd,min have no pole-zero cancellation at γ ∈ Γ . This
completes the proof.
Example 5.2.4. Consider the same G as in Example 5.2.3 with Γ = {1}. For
Gd,min1 in (5.29), one sees
" 1 1 #" # " s−1 #
s+1 s+2 1 − s+1
s+2 s(s+1) 0
GGd,min1 = 1 s s+2
= 1
,
(s−1)(s+1) (s−1)(s+2) − s(s+1) 1 0 s+2
and Gd,min2 at γ = 1 too. Thus, none of Gd,min can meet the condition in
Theorem 5.2.5 and so (PS) for this plant is not solvable. ♦
Example 5.2.5. Let
" s 1 #
s+1 s+2
G(s) = 1 1
.
s(s+1) s(s+2)
Choose
s+1
s−1 − s(s+1)
s−1
Gd = G−1 = .
s+2 s2 (s+2)
− s−1 s−1
It is readily seen that the only common RHP pole and zero of G(s) is 1,
lzγ = 0 and lpγ = 1. From Table 5.3, one can see that Gd,min is not unique.
For
s+1 s(s+1)
s−1 − s−1
Gd,min1 = 2
,
s+2 s (s+2)
− s−1 s−1
we have
" #
s 1 s+1
− s(s+1)
GGd,min1 =
s+1 s+2
s−1 s−1
= 10 ,
1 1 s+2
− s−1 s2 (s+2) 01
s(s+1) s(s+2) s−1
" s(s+1)
#
s+1 − s−1
(1, 0) s2 (s+2)
2
−(s + 2) s−1
" #
s+1
s−1 −s(s − 1)
(0, 1) s+2 2
− s−1 s2 (s + 2)
s + 1 −s(s + 1)
(1, 1) 1
−(s + 2) s2 (s + 2)
1 1
s−1 s(s−1)
0
(−1, −1, 0) 1 1
s 1
s−1 s−1
0 0 1
1 1
− s−1 s 0
(−1, 0, −1) 1
1 s 3
s−1
0 0 1
1 1
s−1 s 0
(−1, 0, 0) 1
1 s 2
s−1
0 0 1
1
1 s(s−1)
0
1 s
(0, −1, −1) 1 s−1 s−1 3
1
0 0 s−1
1
1 s(s−1)
0
(0, −1, 0) 1 1
s 2
s−1
0 0 1
1 1s 0
1 1 s
(0, 0, −1) s−1 2
1
0 0 s−1
1 1s 0
(0, 0, 0) 1 1 s 1
0 0 1
Proposition 5.3.1. Let G ∈ Rm×l sp (s) be a given plant of full row rank and
G = [G+ 0]G− be a stability factorization. Then (P) is solvable if and only if
there exists a K1 ∈ Rm×m (s) such that G+ K1 is block diagonal, nonsingular
and has no pole-zero cancellations, i.e.,
G+ K1 ∈ Rm×m
d (s), (5.34)
and
δ(G+ ) + δ(K1 ) = δ(G+ K1 ). (5.35)
Proof. Assume that K solves (P). Then by Theorem 5.1.1 GK is block diag-
onal and has no RHP pole-zero cancellations. Let Im be the m × m identity
matrix and
K1 := [Im 0]G− K.
Then GK = G+ K1 (which implies (5.34)) and δ + (K) ≥ δ + (K1 ). Moreover,
G+ K1 has no RHP pole-zero cancellations, because
δ + (G+ ) + δ + (K1 ) ≤ δ + (G) + δ + (K) = δ + (GK) = δ + (G+ K1 ).
Since G+ has only RHP zeros and poles and G+ and K1 are both square,
this means that G+ K1 has no zero-pole cancellations at all, which is (5.35).
Conversely, if K1 satisfies (5.34) and (5.35), then
K1
K := (G− )−1 P −1 C (5.36)
0
solves (P), where P ∈ Rm×m
d [s] is stable and such that (G− )−1 [K1T 0]T P −1
m×m
is proper, and C ∈ Rdp (s) internally stabilizes G+ K1 P −1 .
Proposition 5.3.1 simplifies (P) into the following open-loop block de-
coupling problem with stabilizability for a square and nonsingular rational
matrix having only RHP zeros and poles. Note again that K1 is not required
to be proper.
Its pole and zero polynomial is given by (s − 1)(s + 1)2 and (s − 1)2 , respec-
tively. A stability factorization is defined by
1
s 1 0 0 s+1
−1 s
s−1 s−1 s−1 0 1 0 1
+ − s+1
G = −1 1 −s , G = .
0 0 1
0
0 0 s−1 s+1
1
0 0 0 s+1
5.3.1 Solvability
The notion of strict adjoints plays a vital rule in solving the diagonal decou-
pling problem. Its generalization to the block case is obviously required to
solve the general problem (PO).
where P̃i ∈ Rm×mi (s). Let Ai ∈ Rm×mi [s] and Bi ∈ Rmi ×mi [s] be right
coprime such that P̃i = Ai Bi−1 , i = 1, 2, · · · , v. The following lemma shows
that
A := [A1 A2 · · · Av ] (5.37)
Lemma 5.3.1. Any nonsingular polynomial matrix P has a right strict block
adjoint with respect to every partition. A particular right strict block adjoint
of P is given by Prs = A, as defined in (5.37). For a fixed partition, right
strict block adjoints of P are uniquely determined up to multiplication by a
block diagonal unimodular matrix from the right.
s s
s−1
and
−1
0
s−1
−1 0 −1
s−1 0
P̃2 =
1
s−1
0 = 1 0 = A2 B2−1 .
0 1
−s −s 1
s−1 1
Thus a right strict block adjoint Prs (1, 2) with respect to the partition (1, 2)
is given by
s −1 0
Prs (1, 2) = −1 1 0 .
s −s 1
Using the same approach, a right strict block adjoint Prs (2, 1) with respect
to the partition (2, 1) can be computed to be
5.3 Block Decoupling for General Plants 153
1 −1 0
Prs (2, 1) = 0 1 0 .
0 −s 1
We now turn back to (PO) as defined before and choose right coprime
polynomial matrices N + and D+ such that
The matrix N + is nonsingular, and, by Lemma 5.3.1, it has a right strict block
+
adjoint Nrs . For the special case that D+ is the identity matrix (which means
+
that G is stable), K1 = Nrs solves the problem. Using the same approach
for the general case, we have to guarantee that the order of multiplication of
+
Nrs and (D+ )−1 can be appropriately interchanged. Let us assume that Nrs +
which is block diagonal. Let us assume that the above fraction is coprime.
Then there holds
+
δ(G+ K1 ) = deg(det(D̄ls )) + deg(det(D̄+ ))
+
= deg(det(D̄ls )) + deg(det(D+ ))
≥ δ(K1 ) + δ(G+ ).
Hence there are no pole-zero cancellations, and K1 solves (PO). The follow-
ing proposition shows that the assumptions made for the above solution of
the problem are also necessary for solvability of (PO).
Proof. Sufficiency has been shown above. The proof of necessity is more in-
volved, but follows the same lines as the proof of Theorem 5.2.2. At those
places where that proof uses the equality Nd Dd−1 = Dd−1 Nd for diagonal poly-
nomial matrices we have to introduce block diagonal and coprime polynomial
matrices with Nd Dd−1 = D̃d−1 Ñd . The details are omitted.
Propositions 5.3.1 and 5.3.2 directly imply the following solvability con-
dition for the block-decoupling problem with stability using unity output
feedback.
154 5. Polynomial Matrix Approach
Theorem 5.3.1. Let G ∈ Rm×l sp (s) be a given plant of full row rank and
G = [G+ 0]G− be a stability factorization. Then (P) is solvable if and only
+ + +
if Nrs and D+ are externally skew prime and, moreover, N + Nrs
+
and D̄ls D̄
are right coprime.
Let us recapitulate the main steps to check solvability of (P) and to
compute a controller K solving (P): obtain a stability factorization G =
[G+ 0]G− using the Smith-McMillan form of the plant; determine a polyno-
mial fraction as in (5.38), using the Smith-McMillan form of G+ , for instance;
+
construct a right strict block adjoint Nrs of N + as described in Lemma 5.3.1;
check if Nrs and D are externally skew prime and determine D̄+ and N̄rs
+ + +
+ +
satisfying (5.39); determine a left strict block adjoint D̄ls of D̄ as described
+ +
in Lemma 5.3.1; check if N + Nrs +
and D̄ls D̄ are right coprime; determine
K1 = N̄rs (D̄ls ) ; find a stable polynomial matrix P ∈ Rm×m
+ + −1
d [s] such that
(G− )−1 [K1T 0]T P −1 is proper; determine C ∈ Rm×m dp (s) which internally
+ −1
stabilizes the block diagonal G K1 P ; calculate K from (5.36).
Example 5.3.3. Consider the plant of Example 5.3.1. (P) is solvable for G if
and only if (PO) is solvable for
−1
1 0 0 1 1 0
G+ = 0 (s − 1) 0 1 s s = N + (D+ )−1 .
0 0 (s − 1) 0 0 1
Note that G+ has a zero of multiplicity 2 at s = 1 coinciding with a single pole
at s = 1. Since N + is diagonal, we can take Nrs +
= I for any partition. Hence
+ + +
Nrs and D are externally skew prime and N̄rs = I and D̄+ = D+ satisfy
(5.39). In order to determine the solvability of (PO) for this example, we
+ + + +
need only to check if N + Nrs +
= N + and D̄ls D̄ = Dls D are right coprime,
+
where Dls is a left strict block adjoint of D . We have (D+ )T = P , where
+
+ + + T T
P is given in Example 5.3.2, and Dls D = ((D+ )T (Dls ) ) . Hence, for the
partitions (2, 1) and (1, 2), there holds
1 1 0
+
Dls (2, 1)D+ = (Prs (2, 1))T D+ = 0 s − 1 0 ,
0 0 1
and
s−1 0 0
+
Dls (1, 2)D+ = (Prs (1, 2))T D+ = 0 s − 1 0 .
0 0 1
+
One easily sees that N + and Dls (2, 1)D+ are right coprime but N + and
+ +
Dls (1, 2)D are not. Hence, by Theorem 5.3.1, (P) is solvable for the par-
tition (2, 1) but not solvable for the partition (1, 2). For the partition (2, 1),
+
the controller K1 = (Dls (2, 1))−1 solves (PO), and the controller K defined
by (5.36) solves (P). ♦
5.3 Block Decoupling for General Plants 155
stable K̃2 ∈ R(l−m)×m (s) and a polynomial matrix D1 such that N1 D1−1 is a
right coprime polynomial matrix fraction of K1 .
where deg + denotes the degree of the unstable part. Hence, K2 = Ñ2 D̃−1 =
K̃2 D1−1 with K̃2 = Ñ2 D̃2−1 being stable.
It is well known that the achievable performance of a system is essentially
determined by the RHP open-loop poles and zeros. Therefore, in view of our
study in the following section, it is of interest to determine the RHP poles
and zeros which are necessarily introduced by decoupling, i.e., the RHP poles
+ + −1
and zeros of the coprime fraction K̃1 = N̄rs (D̄ls ) .
5.3 Block Decoupling for General Plants 157
+ +
Corollary 5.3.1. The polynomials det(N̄rs ) and det(D̄ls ) divide
+ m−1 + m−1
[det(N (s))] and [det(D (s))] , respectively. Moreover, the controller
+ + −1
K̃1 = N̄rs (D̄ls ) is stable (respectively, minimum phase) if and only if Dg+
(respectively, Ng+ ) as defined in Corollary 5.1.1 is block diagonal.
+ + +
Proof. We have det(N̄rs ) = det(Nrs ), where Nrs is a right strict block ad-
+ +
joint of N . From the equation N (adj(N )) = (det(N + )I it follows that
+
Example 5.3.5. Consider the plant of Example 5.3.3. For the partition (2, 1),
it has been shown that (PO) is solvable. In addition, one can check that G+ U
+
is block diagonal for the unimodular matrix U = (Dls (2, 1))−1 . Moreover, the
controller K1 given in Example 5.3.3 is free of unstable poles and zeros, as
expected. ♦
Using the arguments of Doyle and Stein (1981), we intuitively expect that
no achievable performance is lost by decoupling if Ng+ and Dg+ are both block
diagonal. This conjecture is verified in the following section for a large class
of performance measures.
In this section we assume that the plant G and partition (mi ) are such that
the solvability conditions of Theorem 5.3.1 are satisfied. We study to what
extend the remaining free parameters, as specified in Theorem 5.3.2, can be
used to satisfy additional specifications. We determine a class of plants for
which the achievable performance after decoupling is not inferior to the one
in not necessarily decoupled systems.
It is well known (Freudenberg and Looze, 1988) that many performance
specifications, such as disturbance rejection, tracking and robustness can be
formulated in terms of S = (I +GK)−1 and T = GK(I +GK)−1 . We consider
performance measures J of the form
S
J( ) : R2m×m (s) → R.
T
158 5. Polynomial Matrix Approach
satisfying
S d(S)
J( ) ≥ J( ), (5.41)
T d(T )
and
Td := {T ∈ T : T ∈ Rm×m
d (s)}.
Ng+ ∈ Rm×m
d [s], Dg+ ∈ Rm×m
d [s]. (5.42)
Suppose in addition that det(N + ) and det(D+ ) are coprime. Then Td is equal
to d(T) if and only if (5.42) holds.
5.3 Block Decoupling for General Plants 159
Note that (5.42) holds if and only if there exists a stability factorization
with a block diagonal G+ . Moreover, by Corollary 5.3.1, (5.42) holds if and
only if (PO) can be (easily) solved by the stable and minimum phase con-
troller K1 = (G− )−1 . However, this does not mean that (P) can be solved
by a stable and minimum phase controller, because stabilization might re-
quire unstable poles and zeros in Kd (Vidyasagar, 1985). For the special case
that the plant is already block diagonal, Theorem 5.3.3 means that the cross-
coupling terms in the controller are of no use for performance optimization
with respect to J.
The assumption on the determinants is quite weak; it is satisfied for an
open and dense set of plants. Nevertheless, at least from a theoretical point
of view, it is of interest to analyze if this assumption is really required to
prove the necessity of (5.42). We conjecture that this is not the case.
For a measure J satisfying (5.41), Td = d(T) implies Jd∗ = J ∗ . This
proves the following corollary.
Corollary 5.3.2. Let the plant satisfy (5.42) and the performance measure
satisfy (5.41). Then there are no performance limitations imposed by decou-
pling, in the sense that Jd∗ = J ∗ holds.
Youla, Jabr and Bongiorno (1976) have shown that the optimal controller
of the H2 - problem leads to a decoupled system provided that the plant
is square, stable and minimum phase. Corollary 5.3.2 thus generalizes their
result to possibly unstable, non-minimum-phase, or nonsquare plants, to the
block decoupling problem, and to a wide class of measures including J∞ and
J2 .
Example 5.3.6. Consider the plant of Example 5.3.1. We have already shown
that (P) is solvable if m1 = 2 and m2 = 1. The matrix N + in Example 5.3.3
is already in Hermite form, so that Ng+ = N + . The Hermite form of D+ is
given by
1 1 0
Dg+ = 0 s − 1 0 ,
0 0 1
which is block diagonal with respect to the partition m1 = 2 and m2 = 1.
Thus we will not lose any H∞ - and H2 - performance (in the sense of J∞ and
J2 ) by designing separate optimal controllers Kd,1 and Kd,2 for
−1
+ 1 s−1
G1 =
0 1
and G+
2 = s − 1, respectively, and then using
Kd,1 0
−1
K = (G−
g ) 0 Kd,2
0 0
160 5. Polynomial Matrix Approach
internally stabilizes G+ +
2 . Hence, Kd := block diag{Kd,1 , Kd,2 } stabilizes G ,
− −1 T + −
implying that K̄ := (G ) [I 0] Kd stabilizes G = [G 0]G . Moreover,
we have T̄ := GK̄(I + GK̄)−1 = G+ Kd (I + G+ Kd )−1 = d(T ).
Necessity: Assume first that Ng+ is not block diagonal. Let Ni+ ∈
mi ×m
R [s] be the i-th block-row of Ng+ , Li be the greatest common left divisor
of the columns of Ni+ , and L := block diag{Li }. Then, Ng+ = LÑ + for some
polynomial matrix Ñ + , which can be partitioned as follows:
+
Ñ1
Ñ +
2
Ñ + = . , Ñi+ = [0 Ni Mi ],
..
Ñv+
where
Pv Ñi+ ∈ Rmi ×m [s], Ni ∈ Rmi ×mi [s], and Mi ∈ Rmi ×pi [s]; pi :=
j=i+1 mj . Let k (1 ≤ k < v) be the index such that Mk 6= 0 and Mi = 0
5.3 Block Decoupling for General Plants 161
Kd,1
H d1 d2
H
H
HH
j
r1 z1 y1
-d - - d? - G+ q -
2
– 6
K̃
r2 z2 y2
-d - - d?
- G+ q -
2
– 6
for i > k. Since the degrees of the off-diagonal elements of a row in Ng+ are
strictly smaller than the degree of the corresponding diagonal element, the
matrix Nk has at least one zero s0 ∈ C+ . For this zero there exist y0 ∈ Cmi
and z0 ∈ Cpi such that
y0T Nk (s0 ) = 0,
y0T Mk (s0 ) 6= 0,
[0, y0T , z0T ]Ñ + (s0 ) = 0.
+
Consider now a right strict adjoint Nra of N + . Since Ñ + Nra
+
is block diagonal,
the latter equality implies
Theorem 5.3.2 shows that, for any stabilizing and block decoupling controller
K̄ for G, we have GK̄ = Qd Kd , where Qd = N + Nra +
(D̄+ D̄+ )−1 and Kd
stabilizes Qd . The parameterization of all stabilizing controllers Kd for Qd
(Chapter 3) implies that each element T̄ in Td has the form
T̄ = N + Nra
+
H̄d ,
where H̄d is some block diagonal stable rational matrix. Its k-th diagonal
block T̄k becomes T̄k = Lk P̄k , where P̄k = Ñk+ Nra +
H̄d [0 I 0]T satisfies
T
y0 P̄k (s0 ) = 0.
We now construct a particular stabilizing controller K for G such that the
k-th diagonal block Tk of T = GK(I + GK)−1 is of the form Tk = Lk Pk with
y0T Pk (s0 ) 6= 0. This then implies that d(T ) ∈ d(T) but d(T ) 6∈ Td , which will
complete the proof.
Let G = N D−1 be a right coprime polynomial matrix fraction of G with
N = N + N − for some stable polynomial matrix N − . It follows again from the
162 5. Polynomial Matrix Approach
Definition 5.1.10 . Let G ∈ Rm×l (s) have full row rank, and assume that
there are a full row rank G+ ∈ Rm×l (s) and a nonsingular G− ∈ Rl×l (s)
5.4 A Unified Solution 163
such that
G = G+ G− . (5.43)
It follows from Lemma 5.1.1 that this new stability factorization exists for
any full rank G(s). As a convention, the stability factorization in this section
is always referred to that in Definition 5.1.1 0 above.
For a full row rank polynomial matrix P ∈ Rm×l [s], denote by ∆(P ) the
m-th determinantal, the monic greatest common divisor of all m × m minors
of P (s). In terms of the Smith form, P = U [Λ 0]V with U and V unimodular
and Λ diagonal, there holds ∆(P ) = det(Λ). For a square P , ∆(P ) equals
det(P ) up to a nonzero constant. Let P = P + P − be a stability factorization
as in Definition 5.1.10 , define ∆+ (P ) := ∆(P + ), deg(P ) := deg(∆(P )), where
the right hand side of the last equality means the degree of the polynomial
∆(P ), and deg + (P ) = deg(P + ) = deg(∆(P + )). P is called antistable if
∆(P ) = ∆(P + ), that is, P has all its zeros in C+ .
Let P ∈ Rm×l [s] be a full row rank polynomial matrix. The representa-
tion, P = LU , is said to be an LU-factorization of P if L is a nonsingular
polynomial matrix and U is an irreducible polynomial matrix, i.e., U (s) has
full row rank for all s ∈ C.
Lemma 5.4.1. For an LU-factorization, P = LU , there holds ∆(P ) =
∆(L). In general, let P ∈ Rm×l [s] and P1 ∈ Rm×k [s] be both of full row
rank with k ≥ m such that P = P1 P2 for some polynomial matrix P2 , then
∆(P1 ) divides ∆(P ).
Proof. P has a Smith factorization, P = U1 [Λ 0]U2 , where U1 and U2 are
unimodular with Λ = diag{pi }. By definition, ∆(P ) = ∆(Λ). Express P
as P = LU , where L = U1 Λ and U = [I 0]U2 , then P = LU is an LU-
factorization and ∆(L) = ∆(Λ) = ∆(P ). Now, let P = L̃Ũ be another
LU-factorization, then LU = L̃Ũ . If L−1 L̃ = U3 is unimodular, there will
hold ∆(L̃) = ∆(L) = ∆(P ), which then completes the proof. Let AB −1 be a
coprime polynomial matrix fraction of U3 , then U = AB −1 Ũ . Lemma 5.1.2
says that Ũ has B as its left divisor but Ũ is irreducible and B is thus uni-
modular. Consider dually Ũ = BA−1 U , one concludes that A is unimodular,
too. The second statement follows from applying the Binet-Cauchy formula
(Gantmacher,1959) to the equality, P = L1 P̃ , where P1 = L1 U1 is an LU-
factorization and P̃ = U1 P2 .
A strange thing about nonsquare polynomial matrices is that an anti-
stable P can have a stable divisor for itself. For instance, we have [s 0] =
164 5. Polynomial Matrix Approach
[s 0]diag{1, s + 1}. However, the following lemma shows that such a stable
divisor has certain structure.
Lemma 5.4.2. Let P ∈ Rm×l [s] and P1 ∈ Rm×l [s] be both of full row rank
such that P = P1 P2 for some stable and nonsingular polynomial matrix P2 .
If P is antistable, then U P2−1 is a polynomial matrix, where P = LU is an
LU-factorization.
Proof. Assume conversely that it is not the case, let P̃2−1 Ũ be a left coprime
polynomial matrix fraction of U P2−1 with P̃2 stable and deg(P̃2 ) nonzero,
it then follows that LP̃2−1 Ũ = P1 . By Lemma 5.1.2, LP̃2−1 is a polynomial
matrix. This means that L has a stable divisor P̃2 which is not unimodular,
contradicting the fact that P is antistable.
The matrices D ∈ Rm×m [s] and N ∈ Rm×l [s] are called left coprime in C+ ,
if
Definition 5.4.1. Let a full row rank N ∈ Rm×l [s] and a nonsingular D ∈
Rm×m [s] be given. N and D are said to be externally skew prime in C+ if
there are a full row rank N̄ ∈ Rm×l [s] and a nonsingular D̄ ∈ Rl×l [s] such
that
DN = N̄ D̄ (5.44)
Note that the skew primeness in C+ and ordinary skew primeness (in C) as
defined by Wolovich (1978) differ only in relevant domains, and all tests and
algorithms for ordinary skew prime polynomial matrices in Wolovich (1978)
and Wang,Sun and Zhou (1989) thus apply to the case of skew primeness in
C+ with the obvious modification of replacing C by C+ . Our development
does however require skew primeness in C+ . The following lemma shows that,
under certain conditions, skew primeness in C+ is equivalent to the ordinary
skew primness.
Lemma 5.4.3. Let a full row rank N ∈ Rm×l [s] and a nonsingular D ∈
Rm×m [s] be both antistable, then N and D are externally skew prime in C+
if and only if they are externally skew prime (in C).
5.4 A Unified Solution 165
Proof. Assume that N and D are externally skew prime in C+ , then there are
internally skew prime N̄ and D̄ in C+ . By performing a stability factorization
for D̄T , we get D̄ = D̄− D̄+ with D̄− stable and D̄+ antistable. (5.44) then
becomes
DN = Ñ D̃,
(5.45) and (5.46) together mean that D and Ñ are left coprime. In a similar
way, one can show that N and D̃ are right coprime, too. Thus, it follows that
N and D are externally skew prime. The converse implication is trivial. The
proof is completed.
As one will see later, our solvability condition and compensator parame-
terization for (P) deal only with antistable N and D. Then, due to Lemma
5.4.3, a natural question which one is likely to ask is why the skew primeness
in C+ is here introduced. The real reason is that in spite of the equivalence
there do exist differences between the ordinary internally skew prime pair
(Ñ , D̃) and the internally skew prime pair ( N̄ , D̄) in C+ , as seen in the above
proof and will be further exhibited by the following example. Even more im-
portantly, it will turn out that a general compensator which solves (P) is
constructed from (N̄ , D̄) but not from (Ñ , D̃), and it therefore proves to be
necessary to introduce skew prime polynomial matrices in C+ .
One also sees that D and Ñ are left coprime and N and D̃ are right coprime.
(Ñ , D̃) is thus an internally skew prime pair of (N, D). Further, for a pair
(Ñ , D̃) to be an internally skew prime one of (N, D), it must have certain
properties. Take D̃ for comparison, it satisfies
(5.47) and (5.48) clearly exhibit an essential difference between ordinary skew
prime pairs and skew prime pairs in C+ for the same pair of externally skew
prime polynomial matrices N and D even when they are both antistable. ♦
For an ordinary skew prime pair (Ñ , D̃) of (N, D), we have that ∆(D) =
∆(D̃) and ∆(N ) = ∆(Ñ ) since N D̃−1 = D−1 Ñ with both fractions coprime.
In general, this does not hold for a skew prime pair in C+ . The following
lemma gives a test for skew primeness in C+ and characterizes zero poly-
nomials of an internally skew prime pair in C+ in terms of those of a given
externally skew pair.
Lemma 5.4.4. Let a full row rank N ∈ Rm×l [s] and a nonsingular D ∈
Rm×m [s] be both antistable, then
(i) N and D are externally skew prime in C+ if and only if there are N̄
and D̄ such that (5.44) holds true with ∆+ (D̄) = ∆(D) and N and D̄ right
coprime in C+ ;
(ii) for an internally skew prime pair (N̄ , D̄) in C+ of (N, D), there holds
the equality ∆(N̄ ) = ∆(N ); and
(iii) the equality ∆(D̄) = ∆(D) also holds if N is square. In this case,
there is no difference between ordinary internally skew prime pairs and in-
ternally skew prime pairs in C+ .
Proof. (i): For the sufficiency, we need only to show that D and N̄ are left
coprime in C+ . This is true if there holds
where D1−1 N̄1 is a coprime polynomial matrix fraction of D−1 N̄ . Let N1 (D̄1 )−1
be a coprime polynomial matrix fraction of N (D̄)−1 . As N (D̄)−1 has been
assumed to be coprime in C+ , there holds
(5.44) can be rewritten as N1 (D̄1 )−1 = (D1 )−1 N̄1 with both sides being
coprime, implying that
Collecting (5.50), (5.51), and the assumed condition, deg + (D̄) = deg(D),
gives (5.49).
For necessity, assume now that antistable matrices N and D are externally
skew prime and a dual internally skew prime polynomial matrices in C+ are
N̄ and D̄. Let R be a greatest common right divisor of N and D̄ so that
where N1 D̄1−1 is a right coprime. Moreover, D and N̄ are coprime since they
are coprime in C+ and D is antistable. Therefore, (5.53) means that ( N̄ , D̄1 )
is an internally skew prime pair of (N1 , D) and we thus have
which implies that D̄1 is antistable, too. Since N and D̄ are coprime in C+ ,
their common divisor R must be stable and ∆+ (D̄) = ∆+ (D̄1 R) = ∆+ (D̄1 ) =
∆(D̄1 ). (5.54) then becomes ∆+ (D̄) = ∆(D).
(ii): It follows also from (5.53) that
Lemma 5.4.5. Let P be a full row rank polynomial matrix, then P has a
least right block multiple with respect to every partition. A particular least
right block multiples of P is given by Prm = B, as defined in (5.58). Further,
for a fixed partition, each diagonal block in a least right block multiple of P is
right unimodular equivalent to the corresponding diagonal block in all other
least right block multiples of P .
where Pm = block diag{Pmii }. One sees from the above equation that
Ai Bi−1 Pmii are polynomial matrices, but as Ai Bi−1 are coprime, Bi−1 Pmii
are thus polynomial matrices, or Pm = BPd for some block diagonal polyno-
mial matrix Pd . Therefore, B is a left divisor of any right block multiple of P
0
and by definition it is a least right block multiple of P . Let Prm and Prm be
both least right block multiples of P with respect to the same partition, it fol-
0 0
lows from the definition that Prm = Prm P1 and Prm = Prm P2 for some block
diagonal polynomial matrices P1 and P2 , implying that Prm = Prm P2 P1 and
both P1 and P2 are unimodular. Since they are block diagonal, all the di-
agonal blocks of P1 and P2 are unimodular, too. The proof is completed.
Its inverse is
s −1
s−1 s−1 0
L−1 =
−1
s−1
1
s−1 0
.
s −s
s−1 s−1 1
s −1
s−1 s−1
1 −1 −1
L̃1 = −1 1 = 0 1 1 0
= A1 B1−1 ,
s−1 s−1 1 s−1
s −s 0 −s
s−1 s−1
and
0 0
L̃2 = 0 = 0 [1]−1 = A2 B2−1 .
1 1
We then obtain a least right block multiple Prm (2, 1) with respect to the
partition {2, 1} as
1 0 0
Prm (2, 1) = 1 s − 1 0 .
0 0 1
s s
s−1
and
−1
0
s−1
−1 0 −1
L̃2 = 1
0 = 1 0 s−1 0 = A2 B2−1 .
s−1 0 1
−s −s 1
s−1 1
A least right block multiple Prm (1, 2) with respect to the partition {1, 2} is
given by
s−1 0 0
Prm (1, 2) = 0 s − 1 0 . ♦
0 0 1
We are now interested in the zero property of least right block multiples.
In Example 5.4.3, there hold ∆(P ) = (s − 1), ∆(Prm (2, 1)) = (s − 1), and
∆(Prm (1, 2)) = (s − 1)2 . The following lemma relates the zero polynomial of
Prm to that of P .
Lemma 5.4.6. Let Prm be a least right block multiple of a full row rank poly-
nomial matrix P , then, ∆(P ) divides ∆(Prm ) and ∆(Prm ) divides (∆(P ))m .
5.4 A Unified Solution 171
Proof. By definition, for any least right block multiple Prm there is a poly-
nomial matrix Pra such that P Pra = Prm , which implies from Lemma
5.4.1 that ∆(P ) divides ∆(Prm ). Let P = LU be an LU-factorization and
X a polynomial matrix such that U X = I. Take Pa = X adj(L), where
adj(L) is the adjoint of L, that is, L adj(L) = det(L) I, It then follows
that det(L) I is a right block multiple of P with respect to any partition
as P Pa = LU X adj(L) = det(L) I. Therefore, there is a block polynomial
matrix P1 such that det(L) I = Prm P1 . By Lemma 5.4.1, ∆(Prm ) divides
∆(det(L) · Im ) = (∆(L))m = (∆(P ))m , noting that det(L) differs from ∆(L)
only by a non-zero constant. Hence, the result.
Lemma 5.4.6 shows that Prm does not have any zeros different from those
of P . Especially, Prm will be stable (respectively, antistable) if P has this
property. Example 5.4.3 demonstrates that multiplicity of a zero of Prm may
be greater than the corresponding multiplicity in P .
If the diagonal partition is taken into account, that is, mi = 1, i =
1, 2, · · · , v, v = m, one can expect a natural connection between least right
block multiples and strict right adjoints. The following corollary serves this
purpose.
It is obvious that the concept of least left block multiples for a full column
rank polynomial matrix can be defined in a dual way and the dual results to
least right block multiples all hold.
Let us consider (P) now. Combining Theorem 5.1 with the stability factor-
ization in (5.43), we can obtain the following simplification.
Theorem 5.4.1. For any full row rank plant G ∈ Rm×l p (s), there is a K
such that GK is block diagonal and nonsingular, and has no RHP pole-zero
cancellations if and only if there is a K + which can do so with G replaced by
G+ , where G = G+ G− is a stability factorization.
Proof. For a stability factorization G = G+ G− , we have
δ + (K) = δ + (K + ). (5.60)
if and only if
δ + (G+ K + ) = δ + (G+ ) + δ + (K + ).
G+ = (D+ )−1 N + ,
where D+ ∈ Rm×m [s] is nonsingular, N + ∈ Rm×l [s] has full row rank, and
D+ and N + are both antistable. By a dual statement to Lemma 5.4.5, there
+
is a least left block multiple Dlm ∈ Rm×m
d [s] of D+ such that
5.4 A Unified Solution 173
+ + +
Dla D = Dlm
+
for some nonsingular Dla ∈ Rm×m [s]. It follows from Lemmas 5.4.1 and 5.4.6
+ +
that Dla are antistable. Now, if N + and Dla are externally skew prime in
+
C+ , then there exist internally skew prime N̄ + ∈ Rm×l [s] and D̄la ∈ Rl×l [s]
in C+ such that
+ + +
Dla N = N̄ + D̄la
+
with N̄ + of full row rank and D̄la nonsingular. Let N̄rm+
be a least right
+
block multiple of N̄ . We are now in a position to state the main solvability
condition for (P).
Theorem 5.4.2. For a full row rank plant G, (P) is solvable if and only
+ +
if N + and Dla are externally skew prime in C+ and Dlm +
and N̄rm are left
coprime.
Lemma 5.4.8. Let (D+ )−1 N + and Nk Dk−1 be, respectively, coprime poly-
nomial matrix fractions of G+ and K, G+ K have no unstable pole-zero can-
cellations, and D+ and N + be both antistable, then (D+ )−1 (N + Nk ) is left
coprime and (N + Nk )(Dk )−1 is right coprime in C+ .
Proof. Assume conversely that (D+ )−1 (N + Nk ) is not coprime, then a coprime
one (D̃+ )−1 Ñ of it will satisfy
We therefore have
δ + (G+ K) = δ + ((D̃+ )−1 Ñ Dk−1 )
≤ deg (D̃+ ) + deg + (Dk )
< deg (D+ ) + deg + (Dk )
= δ + (G+ ) + δ + (K),
which contradicts the assumption that G+ K has no unstable pole-zero cancel-
lations. (D+ )−1 (N + Nk ) is thus left coprime. The coprimeness of (N + Nk )(Dk )−1
in C+ follows similar lines. The proof is completed.
174 5. Polynomial Matrix Approach
If K solves (P) for G+ , it follows from Theorem 5.4.1 that K is such that
G+ K is block diagonal and has no unstable pole-zero cancellations. The latter
fact further implies by Lemma 5.4.8 that (D+ )−1 (N + Nk ) is coprime and
Lemma 5.4.7 then says that
Dq = D1 D + (5.63)
where
−1 −1
N̄q D̄q1 = Dq1 Nq (5.66)
is block diagonal and polynomial matrices N̄q and D̄q1 are chosen such that
−1
they are both block diagonal and right coprime. Since N̄q D̄q1 is coprime and
(5.65) holds, D̄q1 is a left divisor of Dk , i.e.,
for some polynomial matrix Dk1 . Substituting (5.67) into (5.65) yields
+ + −1
Dla N D̄k1 N̄k = N̄q , (5.68)
where
−1 −1
D̄k1 N̄k = Nk Dk1 (5.69)
with the left hand side of (5.69) being a left coprime polynomial matrix
fraction. Note from (5.68) that this coprimeness, together with the fact that
N̄q is a polynomial matrix, implies, by Lemma 5.1.2, that
5.4 A Unified Solution 175
+ + −1
Dla N D̄k1 = N̄ + (5.70)
is a polynomial matrix and of full row rank. One knows from Lemma 5.4.8
that no unstable pole-zero cancellations in G+ K implies that N + Nk and Dk
are right coprime in C+ and this also ensures that
and
Since N̄q is block diagonal, it follows from Lemma 5.4.5 that there is some
N̄q1 ∈ Rm×m
d [s] such that
+
N̄rm N̄q1 = N̄q , (5.80)
176 5. Polynomial Matrix Approach
+
where N̄rm is a least right block multiple of N̄ + . From (5.64), (5.66), and
(5.80), Q = G+ K can be expressed as
+ −1 + −1 −1
Q = Dq−1 Nq = (Dq1 Dlm ) Nq = (Dlm ) Dq1 Nq
+ −1 −1 + −1 + −1
= (Dlm ) N̄q D̄q1 = (Dlm ) N̄rm N̄q1 D̄q1 . (5.81)
+ −1 +
Because Dq−1 Nq is coprime, so is (Dlm ) N̄rm . The proof is completed.
SUFFICIENCY: If the conditions in the theorem are satisfied, there is an
+
internally skew prime pair (N̄ + , D̄la ) in C+ such that
+ + +
Dla N = N̄ + D̄la (5.82)
+
with N + and D̄la right coprime in C+ , and
+ +
∆(Dla ) = ∆+ (D̄la ). (5.83)
+
Let N̄rm be a least right block multiple of N̄ + such that
+
N̄ + N̄ra +
= N̄rm (5.84)
+
for some polynomial matrix N̄ra . We then take
+ −1 +
K1 = (D̄la ) N̄ra . (5.85)
as a candidate for the required compensator K. One sees that
δ + (G) = deg(D+ ), (5.86)
and
+
δ + (K1 ) ≤ deg + (D̄la ). (5.87)
By (5.82) and (5.84), the loop transfer matrix can be rewritten as
+ −1 +
Q1 = GK1 = (D+ )−1 N + (D̄la ) N̄ra
+ −1 + + + + −1 + +
= (D+ )−1 (Dla ) N̄ N̄ra = (Dla D ) N̄ N̄ra
+ −1 +
= (Dlm ) N̄rm , (5.88)
+
which is block decoupled. The assumed left coprimeness of Dlm and N̄rm
implies that
+ +
δ + (Q1 ) = deg(Dlm ) = deg(D+ ) + deg(Dla )
+ + +
= deg(D ) + deg (D̄la ), (5.89)
where the last equality follows from (5.83). Therefore, collecting (5.86), (5.87),
and (5.89) gives
δ + (Q1 ) ≥ δ + (G) + δ + (K1 ), (5.90)
which ensures that no unstable pole-zero cancellations will occur in GK1 .
Thus, by Theorem 5.4.1, (P) is solvable for G+ . The sufficiency is proven.
Let us take an example for illustration.
5.4 A Unified Solution 177
It has been shown in Section 5.2 that a sufficient condition for solvability
of the diagonal decoupling problem with stability for a square plant is that
the plant has no unstable poles coinciding with zeros and plants generically
satisfy this condition. The same result is established for the general block
decoupling case with arbitrary plants as follows.
Corollary 5.4.2. If a full row rank plant G has no unstable poles coinciding
with zeros, then (P) is solvable for every partition.
+ + + + +
Proof. It follows from Dla D = Dlm that ∆(Dla ) divides ∆(Dlm ) and it
+ + m
further implies, by Lemma 5.4.6, that ∆(Dlm ) divides (∆(D )) . Therefore,
+
under the assumed condition, N + and Dla have no zeros in common and they
are externally skew prime in C and so are in C+ . It follows from Lemmas 5.4.6
+
and 5.4.4 that ∆(Dlm ) divides (∆(D+ ))m and ∆(N̄rm +
) divides (∆(N̄ + ))m =
+ m + +
(∆(N )) , so that square polynomial matrices Dlm and N̄rm have no zeros
in common and they are thus left coprime. Hence, by Theorem 5.4.2, (P) is
solvable. The proof is completed.
In fact, for the case considered in the above corollary, we can directly
construct a decoupling compensator as follows. Let N + = L+ U be an LU-
factorization and X a polynomial matrix such that U X = I, then take K
as
so that GK is block decoupled for any partition. Since ∆(D+ ) and ∆(N + )
have been assumed to be coprime in C+ , there are no unstable pole-zero
cancellations in GK. The disadvantage with this decoupling compensator is
that its order may be much higher than necessary. In general, lower order ones
can be obtained from the parameterization of all decoupling compensators,
which will be presented in the next section.
In the extreme case of one block partition, i.e., v = 1, m1 = m, our (P)
defects to nothing than stabilization and the conditions in Theorem 5.4.2
+
should hold automatically for any plant G. Indeed, in this case, Dlm = D+ ,
+ + + +
Dla = I, N̄rm = N̄ = N . The solvability conditions are then reduced to
ones that N + and I are externally skew prime in C+ and D+ and N + are
left coprime, which, of course are satisfied for any G+ .
Proof. In the proof of necessity of Theorem 5.4.2, one sees from (5.67), (5.69),
(5.70) and (5.82) that every K which block-decouples G+ and maintains
internal stabilizability cab be expressed as
K = Nk Dk−1 = Nk (D̄q1 Dk1 )−1
−1 −1 −1 −1
= Nk Dk1 D̄q1 = D̄k1 N̄k D̄q1
+ −1 −1
= (D̄la ) N̄k D̄q1 . (5.92)
It follows from (5.79) and (5.80) that N̄k satisfies
N̄ + N̄k = N̄rm
+
N̄q1 . (5.93)
+
Since N̄ra N̄q1 is obviously a particular solution of (5.93) for N̄k and ZW is
a general solution of N̄ + N̄k = 0, a general N̄k which satisfies (5.93) is then
given by
+
N̄k = N̄ra N̄q1 + ZW, (5.94)
Theorem 5.4.3. Assume that a full row rank plant G and a partition be
+ + +
given, the notations, G+ , G− , D+ , N + , Dlm , Dla , N̄ + , D̄la +
, and N̄rm are de-
fined as before, and the conditions in Theorem 5.4.2 are satisfied. In addition,
+
let N̄ra +
be a polynomial matrix such that N̄ + N̄ra = N̄rm +
and Z be an l×(l−m)
minimal polynomial base of the right null space of N̄ + , then
(i) all compensators which block-decouple G with internal stabilizability
are given by
+ −1 −1
K = (G− )−1 (D̄la +
) (N̄ra Nk2 + ZW )Dk2 ,
where polynomial matrices Nk2 and Dk2 are both block diagonal with respect
−1
to the partition and right coprime, K2 = Nk2 Dk2 is coprime and has no pole-
+ −1 +
zero cancellations with Q1 = (Dlm ) N̄rm and, W is an arbitrary polynomial
matrix of size (l − m) × m;
(ii) all achievable open-loop transfer function matrices Q = GK under
block decoupling with internal stabilizability can be expressed as
Q = Q 1 K2 ,
and
+
∆(N̄rm ) = ∆(N + ) (5.97)
Proof. (i) ⇒ (ii): If (P) is solvable and both (5.96) and (5.97) hold true,
+ +
then ∆(Dla ) = 1 and Dla is unimodular. The C+ -skew primeness equa-
+ + + + +
tion, Dla N = N̄ D̄la , can be thus rewritten as N + = P D̄la , where
+ −1 + +
P = (Dla ) N̄ is a polynomial matrix. Note that N is antistable, and
+ + +
D̄la is stable since, by Lemma 5.4.4, ∆+ (D̄la ) = ∆(Dla ) = 1. It then
+ −1
follows from Lemma 5.4.2 that U1 (D̄la ) = P1 is a polynomial matrix,
where N + = L+ U1 is an LU-factorization. Consider now the equality,
+ −1 + + −1 +
(D+ )−1 N + (D̄la ) N̄ra = (Dlm ) N̄rm . It can be rewritten as
+ −1 +
+
(D+ )−1 L+ P1 N̄ra = (Dlm ) N̄rm . (5.98)
+
Using (5.96) and (5.97) with ∆(L+ ) = ∆(N + ) and the coprimeness of Dlm
+ +
and N̄rm , one concludes that U2 = P1 N̄ra is a unimodular polynomial ma-
trix. Take U as U = U2−1 U1 , then U is an irreducible polynomial matrix.
Postmultiplying (5.98) by U yields
+ −1 +
(Dlm ) N̄rm U = (D+ )−1 L+ U1 = (D+ )−1 N + = G+ . (5.99)
+ −1
In view of (5.96) and (5.99) with D+ and N + left coprime, (Dlm +
) (N̄rm U)
is thus left coprime, too.
(ii) ⇒ (iii): Let X be a polynomial matrix such that U X = I, then X
has neither zeros nor poles and K = (G− )−1 X is stable and minimum phase.
The resulting loop transfer matrix becomes
+ −1 + + −1 +
Q = GK = (Dlm ) N̄rm U G− (G− )−1 X = (Dlm ) N̄rm ,
+
∆(N̄rm ) divides ∆(N + ). But by Lemmas 5.4.4 and 5.4.6, ∆(N + ) divides
+
∆(N̄rm ), too. Hence, (5.97) holds true. (5.96) can be proven in a similar way
by noting from (5.100) that ∆+ (Dq ) divides ∆(D+ )∆+ (Dk ) = ∆(D+ ) as K
is stable. The proof is completed.
Example 5.4.5. Consider again the plant in Example 5.4.4. For the partition
{2, 1}, Example 5.4.4 has shown that (P) is solvable. In addition, one easily
+ +
sees that ∆(N̄rm ) = (s − 1) = ∆(N + ) and ∆(Dlm ) = ∆(D+ ) = (s − 1)2 .
Therefore, the conditions in (i) of Theorem 5.4.4 are satisfied. G+ can be
expressed as
+ −1 +
G+ = (Dlm ) N̄rm U =
−1
1 0 0 1 0 0 1 1 0 2
0 (s − 1) 0 1 s − 1 0 0 1 0 1 ,
0 0 (s − 1) 0 0 1 0 s 1 s+1
which is block decoupled with respect to the partition {2, 1}. One also sees
that N + = Nrm +
U and N + Nk = Nrm +
so that ∆(N + ) = ∆+ (N + Nk ) =
+
∆(Nrm ). Thus, (iii) of Theorem 5.4.4 also holds, indeed. ♦
It should be pointed out that (i) or (ii) of Theorem 5.4.4 implies that
there is a stable and minimum phase compensator K such that GK is block
decoupled and has no unstable pole-zero cancellations, but the converse is
not true in general. For instance, consider the plant:
s 1
s−1 s−1 0
G=
0 1 0
−1
s−1 0 s 1 0
=
0 1 0 1 0
= (D+ )−1 N + ,
for the partition {1, 1}. Take K as
5.4 A Unified Solution 183
1 −1 −1
1 0
K= 0 s = Nk Dk−1 ,
0 s+1
0 1
which is stable and minimum phase. One can verify that GK is decoupled
and has no unstable pole-zero cancellations. Simple calculations show that
∆(N + ) = s but ∆+ (N + Nk ) = s2 so that ∆(N + ) 6= ∆+ (N + Nk ) for this K.
Are there any other decoupling K which makes the equality
∆(N + ) = ∆+ (N + Nk ) (5.102)
hold? The answer is No! To see this, one notes from Corollary 5.4.2 that (P) is
solvable for the given plant and the partition {1, 1}, and (5.101) thus applies
+ +
with N̄rm = Nrm = diag{s, s}. It follows that deg + (N + Nk ) ≥ deg + (Nq ) ≥
deg (Nrm ) = 2 > 1 = deg(N + ). Therefore, among all the compensators K
+ +
Theorem 5.4.5. Given a full row rank plant G ∈ Rm×l p (s) and a partition
{mi }, let D−1 N = G be a left coprime polynomial matrix fraction and Dlm
a least left block multiple of D with respect to the partition such that Dlm =
Dla D for some polynomial matrix Dla . If N and Dla are externally skew
prime, construct internally skew prime polynomial matrices Ñ and D˜la such
that Dla N = Ñ D˜la . And let Ñrm be a least right block multiple of Ñ with
respect to the partition such that Ñ N˜ra = N˜rm for some polynomial matrix
N˜ra and Z a l × (l − m) minimal polynomial base of the right null space of
Ñ , then the following hold true:
(i) there is a proper compensator K in a unity output feedback system
such that GK is nonsingular and block diagonal with respect to the partition
184 5. Polynomial Matrix Approach
where polynomial matrices Nk2 and Dk2 are both block diagonal with respect
to the partition and right coprime, Q1 K2 has no pole-zero cancellations with
−1 −1
Q1 = Dlm Ñrm and K2 = Nk2 Dk2 , and W is an arbitrary polynomial matrix
of size (l − m) × m; And
(iii) all achievable open-loop maps under block decoupling with arbitrary
pole assignability can be expressed as
Q = Q 1 K2 ,
In this section, the block decoupling problem with internal stability for
unity output feedback systems has been considered in a general setting and
completely solved. Necessary and sufficient conditions for its solvability are
given and are only concerned with the greatest C + -divisor of a given plant.
The characterization of decoupled systems is presented and it shows that a
general compensator which solves the problem consists mainly of two parts,
one being the decoupling part which makes the plant block decoupled with-
out unstable pole-zero cancellations, the other the stabilizing part which is
block diagonal and makes the block decoupled plant internally stable and
the whole compensator proper. The block decoupled plant (the plant plus
the decoupling part) need not introduce different zeros or poles from those
of the plant itself and what may happen is that zero or pole multiplicities of
the former are greater than those of the latter. Two new concepts, least block
multiples of a full rank polynomial matrix and skew prime polynomial matri-
ces in C+ , are introduced and they have played important roles, respectively,
in the problem solvability and compensator parameterization.
poles. They show that the diagonal decoupling problem is solvable under this
assumption and provide a parameterization of controllers solving the prob-
lem This assumption is relaxed and the diagonal decoupling problem with
stability is solved by Linnemann and Maier (1990) for 2 × 2 plants, and by
Wang (1992) for square plants, along with a controller parameterization. For
the same problem, alternative necessary and sufficient conditions are also de-
rived by Lin (1997) based on transfer matrices and residues, and in Wang
and Yang (2002) using the minimal C + -decoupler of the plant. The results
of Wang (1992) are generalized to the block decoupling problem for possibly
non-square plants by Linnemann and Wang (1993) with additional discus-
sion on the performance limitations in the decoupled system compared with
the performance which is achievable by not necessarily decoupling controllers.
A unified and independent solution for the general case is also presented
in Wang (1993). The main tools used in this chapter are skew primeness
(Wolovich, 1978), strict adjoints (Hammer and Khargonekar, 1984), stability
factorizations (Wang, 1992), and their extensions to non-square plants and
block diagonalization.
6. Transfer Matrix Approach
In the preceding two chapters, the plants which we have considered assume
no time delay and for such a kind of plants the complete solution has been
developed in both state-space and polynomial matrix settings. However, time
delay often exists in industrial processes and other systems, some of which will
be highlighted in Section 1. We need an entirely new approach to decoupling
problem for plants with time delays. Time delay imposes a serious difficulty
in theoretical development and thus the theory of time delay systems is much
less matured than delay-free systems. In the present and next chapters, we
will address decoupling problem for plants with time delay and our emphasis
is to develop decoupling methodologies with necessary theoretical supports
as well as the controller design details for possible practical applications.
This chapter will consider conventional unity output feedback configuration
whereas the next chapter will exclusively deal with time delay compensation
schemes.
In this chapter, several typical plants with time delay will be briefly de-
scribed and their transfer matrices with time delay given in Section 1 as the
motivation of subsequent discussions on delay systems. The new decoupling
equations are derived in a transfer function matrix setting for delay plants
(Section 2), and achievable performance after decoupling is analyzed (Section
3) where characterization of the unavoidable time delays and non-minimum
phase zeros that are inherent in a feedback loop is given. In Section 4, a
design method is presented. The objective is to design a controller of lowest
complexity which can achieve fastest loops with acceptable overshoot and
minimum loop interactions. Stability and robustness are analyzed in Section
5. Section 6 shows why multivariable PID control often fails when the number
of inputs/outputs is getting bigger. More simulation is given in Section 7 and
the conclusion is drawn in Section 8.
processes are introduced in this section to show the presence of time delay
and will be used in simulation later to substantiate the effectiveness of the
proposed decoupling method for high performance control.
Example 6.1.2. Doukas and Luyben (1978) studied the dynamics of a dis-
tillation column producing a liquid sidestream product, with the objective
of maintaining four composition specifications on the three product streams.
The Tyreus process in (6.1) is a 3 × 3 subsystem of it. The transfer function
matrix for the 4 × 4 model is
−11.3e−3.79s 0.374e−7.75s −9.811e−1.59s −2.37e−27.33s
(21.74s+1)2 22.2s+1 11.36s+1 33.3s+1
5.24e−60s −1.986e−0.71s 5.984e−2.24s 0.422e−8.72s
400s+1 66.67s+1 14.29s+1 (250s+1)2 ;
G(s) =
−0.33e−0.68s
,
(6.2)
0.0204e−0.59s 2.38e−0.42s
(2.38s+1)2 0.513e−s
(7.14s+1)2 (1.43s+1)2
4.48e−0.52s −0.176e−0.48s −11.67e−1.91s
11.11s+1 (6.90s+1)2 12.19s+1 15.54e−s
Example 6.1.3. Alatiqi and Luyben (1986) presented the results of a quanti-
tative study of the dynamics of two alternative distillation systems for sep-
arating ternary mixtures that contain small amounts (less than 20%) of the
intermediate component in the feed. The process transfer function matrix of
the complex sidestream column/stripper distillation process is given by
6.1 Plants with Time Delay 189
where the four controlled variables are y1 (mole fraction of the first component
in the distillate); y2 (mole fraction of the third component in the bottom); y3
(mole fraction of the second component in the sidestream); y4 (temperature
difference); and the four manipulated variables are u1 (reflux flow rate); u2
(main column reboiler heat-transfer rate); u3 (stripper reboiler heat-transfer
rate); and u4 (liquid draw rate from main column to stripper). Although the
complex sidestream column/stripper configuration is more energy efficient
than the conventional two-column sequential ”light-out-first” configuration,
it presents a challenging 4 × 4 multivariable control problem. ♦
Example 6.1.4. Ammonia is commercially produced by synthesis of the raw
material gas (Hydrogen, H2 , Nitrogen, N2 ) via a reversible reaction catalyt-
ically with great energy consumption for making raw gases. The recycled
un-reacted raw gas makes the whole ammonia process more difficult to oper-
ate and control. In general, ammonia process consists of two big sessions. One
is reformer session to produce reactant gas (H 2 , N2 ). Another is conversion
session to synthesize ammonia. Many disturbances and frequent change of
feed gas rate affect a modern ammonia plant. Moreover, these disturbances
always propagate throughout the entire plant. It has been common require-
ments to improve operating stability, increase material and energy efficiency
and prevent disturbances from echoing throughout the plant. A transfer func-
tion matrix of the reformer session was identified as
−4s −s −4s
−0.562e −0.01e 0.378e
8s+1 3s+1 8s+1
−19s
−0.0159e−21s ,
G(s) = 0.0135e
5.6s+1 0 10s+1 (6.4)
−10s −4s −7.5s
0.002e −0.00125e 0.12e
7s+1 3s+1 10s+1
where the controlled variables and manipulated variables are y1 (the primary
reformer coil outlet temperature); y2 (the secondary reformer methane leak-
age) and y3 (the H2 /N2 ratio); u1 (the process natural gas flow rate); u2 (the
process air flow rate); and u3 (the fuel natural gas flow rate). ♦
Example 6.1.5. The purpose of a depropanizer column is to separate propane
from the feed which comes from deethanizer column. The operation target of
this unit is to keep the top butane concentration low than 5% and bottom
propane concentration low than 5%. The main controlled variables includes
y1 (top butane concentration), y2 (bottom propane concentration) and y3
190 6. Transfer Matrix Approach
6.2 Decoupling
Consider the conventional unity feedback control system in Figure 6.1, where
G represents the transfer matrix of the plant and K the multivariable con-
troller. Here, G is assumed to be square and nonsingular:
g11 (s) · · · g1m (s)
.. ,
G(s) = ... ..
. . (6.6)
gm1 (s) · · · gmm (s)
where
and gij0 (s) are strictly proper, stable scalar rational functions and Lij are
non-negative constants, representing time delays. The controller K(s) is an
m × m full-cross coupled multivariable transfer matrix:
k11 (s) · · · k1m (s)
.. ,
K(s) = ... ..
. . (6.7)
km1 (s) · · · kmm (s)
where
θij are non-negative constants, and kij0 (s) are scalar proper rational func-
tions. kij0 (s) should be stable except a pole at s = 0, and for simplicity, it is
called stable in this chapter (referred to the controller only). This pole is de-
sired to have integral control. Assume that the plant has no zero at the origin.
Then, the integral control will lead to zero steady state errors in response to
step inputs provided that the closed-loop is stable. It is also assumed that a
proper input-output pairing has been achieved in G(s) such that none of the
m principal minors of G(s) (the i th principle minor is the determinant of a
(m − 1) × (m − 1) matrix obtained by deleting the i th row and i th column
of G(s)) are zero.
Our task here is to find a K(s) such that the closed loop transfer matrix
from the reference vector r to the output vector y:
−1
H(s) = G(s)K(s) [I + G(s)K(s)] , (6.8)
is decoupled, that is, H(s) is diagonal and nonsingular. One sees that
H −1 = [I + GK] (GK)−1
= (GK)−1 + I, (6.9)
and the closed-loop H(s) is decoupled if and only if the open loop G(s)K(s)
is decoupled, i.e.,
g1
g2
GK = . [k1 k2 · · · km ] = diag{qii }, i = 1, 2, . . . , m. (6.10)
..
gm
192 6. Transfer Matrix Approach
One notes that for a given i, the resulting diagonal element of GK is indepen-
dent of the controller off-diagonal elements but contains only the controller
diagonal element kii .
To further simplify the above result, let Gij be the cofactor corresponding
to gij in G. It follows from linear algebra (Noble, 1969) that the inverse of G
ji
can be given as G−1 = adj G
|G| , where adj G = G . This also means
11
g11 · · · g1m G · · · Gm1
.. . . .. .. .. .
. . . . . .. = |G|Im . (6.17)
gm1 · · · gmm G1m · · · Gmm
Equation (6.17) can be equivalently put into the following two relations:
i1
g11 · · · g1,i−1 g1,i+1 · · · g1m G g1i
.. .. .. .. .. .. .. ..
. . . . . . . .
i,i−1
gi−1,1 · · · gi−1,i−1 gi−1,i+1 · · · gi−1,m G
i,i+1 = − gi−1,i Gii , ∀ i ∈ m,
gi+1,1 · · · gi+1,i−1 gi+1,i+1 · · · gi+1,m G gi+1,i
. .. .. .. .. .. .. .
.. . . . . . . .
.
gm1 · · · gm,i−1 gm,i+1 · · · gmm Gim gmi
(6.18)
and
m
X
gik Gik = |G|, ∀i ∈ m.
k=1
or
Gij
ψji = , ∀ i, j ∈ m, j 6= i. (6.19)
Gii
By (6.19), (6.14) and (6.16) become respectively
Gij
kji = kii , ∀ i, j ∈ m, j 6= i, (6.20)
Gii
194 6. Transfer Matrix Approach
and
m
X m
Gik 1 X |G|
g̃ii = gii + gik ii
= ii
gik Gik = ii , ∀ i ∈ m. (6.21)
G G G
k=1 k=1
k6=i
e−2s (s − 0.5)e−3s
G22 = , G12 = − .
s+2 (s + 2)2
It follows from (6.21) that the decoupled loops have their equivalent processes
as
2(s + 2)e−s + (s − 0.5)e−2s
g̃11 = ,
(s − 0.5)(s + 2)
2(s + 2) 5s
ψ21 = − e , ψ12 = e−4s .
s − 0.5
6.3 Achievable Performance 195
Notice that time delays and non-minimum phase zeros are inherent charac-
teristic of a process and can not be altered by any feedback control. They
Q ni (z)
are reflected in (6.27) respectively by e−Li s and z∈Zi z−s z+s . The fac-
tor of the standard 2nd-order transfer function represents usual performance
requirement such as overshoot and settling time for the minimum phase and
assignable part of the loop. The term ( Ni1ωni s + 1)νi is to provide necessary
high frequency roll-off rate required by decoupling and properness of the
controller.
Firstly, consider the time delay Li of the ith loop. As time delay imposes
performance limitation in a control loop, to achieve best performance and
to make the closed-loop response fastest, the time delay Li in the ith loop
in (6.27) should be chosen as the smallest possible value. If only (6.22) is
concerned, it seems that Li can be chosen as the time delay in g̃ii as kii cannot
contain pure prediction for realizability. However, decoupling may require this
delay to be increased. To see this, assume that the system interaction from
ith input to jth output, gji , has smaller time delay than that of the transfer
function from the jth input to the jth output. Then, complete elimination
of the interaction, gji , by a realizable kji is impossible, unless the ith loop
delay is increased by a certain amount.
It is thus interesting to know when decoupling requires further delay in
a certain loop and by what amount. To address this problem, consider the
process G(s) given by (6.6). A general expression for non-zero Gij and |G|
will be
M
X
φ(s) = φl (s)e−αl s , αl > 0, (6.28)
l=1
It is easy to verify τ (φ1 φ2 ) = τ (φ1 )+τ (φ2 ) for any non-zero φ1 (s) and φ2 (s).
And it is obvious that for any realizable and non-zero φ(s), τ (φ(s)) can not
be negative. It follows from (6.26) that
IDEAL
|G|kji = Gij qrii , ∀ i, j ∈ m. (6.29)
where
Therefore, the time delay Li in the ith objective closed-loop transfer function
is chosen as the minimum among all meeting (6.30):
Li = τ (|G|) − τi . (6.32)
We have just seen how the various time delays in Gij and |G| limit the
control system performance and affect the choice of the objective loop transfer
function. In addition to time delays, zeros of Gij and |G| which lie in the
closed right half plane (RHP) may also limit the control system performance.
To address this problem, let Z+ 1
|G| be the set of RHP zeros of |G|. For a stable
and proper kji , |kji (z)| is finite for all Re(z) > 0. Thus, the left side of (6.29)
will evaluate to 0 for each z ∈ Z+ ij
|G| , and hence, the right side of (6.29), G qrii
+
must also interpolate to 0 for each z ∈ Z|G| . This implies that qrii needs to
have a RHP zero at z if z is itself not a RHP zero of Gij . The situation may
become more complicated when RHP zeros of |G| coincide with those of Gij
and when multiplicity is considered.
In order to investigate what RHP zero must be included in qrii , for a
non-zero transfer function φ(s), let η z (φ) be a non-negative integer ν such
that lims→z φ(s)/(s − z)ν exists and is non-zero. Thus, φ(s) has η z (φ) zeros
at s = z if η z (φ) > 0, or no zeros if η z (φ) = 0. It is easy to verify that
η z (φ1 φ2 ) = η z (φ1 ) + η z (φ2 ) for any non-zero transfer functions φ1 (s) and
IDEAL
φ2 (s). It then follows from (6.29) that a stable and proper kji (s) requires
1
It is assumed in this chapter that |G| has no zero at the origin so as to enable
the inclusion of integral control in K(s) for robust tracking and regulation with
respect to step inputs.
198 6. Transfer Matrix Approach
or
where
Equation (6.33) implies that the ith decoupled open loop transfer func-
tion qrii must contain the RHP zeros at each Z+ |G| with multiplicity of
η z (|G|) − ηi (z) if η z (|G|) − ηi (z) > 0. Note that a feedback connection of
a loop transfer function alter neither the existence nor the location of a zero.
Thus the ith decoupled closed-loop transfer function hri should also contain
these RHP zeros with exactly the same multiplicity. Note also from (6.29)
and the assumed stability of G that qrii need not include any other RHP zero
except those of |G|; and for best performance of the closed-loop system it is
undesirable to include any other RHP zeros in qrii (s) more than necessary.
Therefore, Zi and ni (z) in (6.27) are chosen respectively as
Zi = Z+
|G| , (6.35)
and
We now consider the term ( Ni1ωni s + 1)νi , which provides necessary high
frequency roll-off rate required by decoupling and properness of the controller.
Ni is usually chosen as 10 ∼ 20. To determine the integer νi , consider a
non-zero transfer function φ(s), let ν(φ) be the smallest integer ν such that
limω→∞ |φ(jω)/(jω)(1+ν) | = 0. A transfer function φ is said to be proper if
ν(φ) ≥ 0. It is easy to verify that ν(φ1 φ2 ) = ν(φ1 ) + ν(φ2 ) for any non-zero
transfer functions φ1 (s) and φ2 (s). For realizability, all controller elements
IDEAL
should be proper, i.e., ν(kij ) ≥ 0. It then follows from (6.29) that
or
Since the feedback connection of a loop transfer function does not alter its
relative degree, i.e., ν(qrii ) = ν(hri ) = 2 + νi , (6.37) is equivalent to
and taking
6.3 Achievable Performance 199
ensures that the resultant kij are all proper and thus physically realizable.
Now all the necessary RHP zeros z, time delays Li as well as νi , required
by (6.27), have been determined. They are all feedback invariant. What re-
mains to fix are the parameters ωni and ξi in (6.27). ξi determines the reso-
nant peak magnitude Mri of the frequency response of the closed-loop transfer
function hri . The magnitude of the resonant peak gives an useful indication
of the relative stability of the system (Ogata, 1997). A large resonant peak
magnitude indicates the presence of a pair of dominant closed-loop poles with
small damping ratio, which will yield an undesirable transient response. A
small resonant peak magnitude, on the other hand, indicates the absence of
a pair of dominant closed-loop poles with small damping ratio, meaning that
the system is well damped. By ignoring the term ( 1 1s+1)νi in the operat-
Ni ωni
ing bandwidth, the magnitude of the ith closed-loop transfer function hri in
(6.27) can be calculated as
1
|hri (jω)| = q , (6.39)
ω2
(1 − 2
ωni
)2 + (2ξi ωωni )2
and is given by
rq
ωbi ≈ ωni (2ξi2 − 1)2 + 1 − (2ξi2 − 1)
for hri (s) in (6.27). The loop bandwidth ωbi is usually close to the 0 dB gain
cross-over frequency ωgi , at which |qrii (jωgi )| = 1. Maciejowski (1989) shows
that for typical, acceptable designs we can estimate the loop bandwidth in
terms of the cross-over frequency by
200 6. Transfer Matrix Approach
or
ü ü
ü det(G) ü
|hri | ≤ min{üü ü umax,j }.
j∈Ji Gij ü
1 |G(jωni )|
≤ min{| ij |umax,j }. (6.44)
2ξi j∈Ji G (jωni )
A straightforward method to determine the achievable ωni for ith loop under
the given input constraints is simply to plot the right hand side of (6.44).
1 X üü ü2
N
ü
e= üW (jωk )(φ̂(jωk ) − φ(jωk ))ü , (6.46)
N
k=1
is minimized, where W (jωk ) serves as the weightings. Two very effective al-
gorithms with stability preservation have been presented in Chapter 3, one
using recursive least squares and another using the combination of the FFT
and step response. In what follows in this section, we refer “the model reduc-
tion algorithm” to one of them, or anyone else which can produce a reliable
reduced-order model in the form of (6.45).
To find the objective loop transfer function numerically, we apply the
model reduction algorithm to |G| and Gij to obtain the corresponding
reduced-order models (here, the weighting can be chosen as unity, i.e.,
W (jω) = 1). To get a model with lowest order and yet achieve a speci-
fied approximation accuracy, we start the model reduction algorithm with a
model of order 2 and then increase the order gradually to find the smallest
integer n such that the corresponding maximum relative error is less than a
user-specified (usually 5% ∼ 15%), i.e.,
ü ü
ü φ̂(jω ) − φ(jω ) ü
ü k k ü
max ü ü ≤ . (6.47)
k=1,··· ,N ü φ(jωk ) ü
Algorithm 6.4.1 Determine the objective loop performances qrii (s) given
G(s), Mri , φmi and umax,i .
(i) Apply the model reduction algorithm to |G| and Gij , i, j ∈ m, to obtain
the corresponding reduced-order models. Determine the set Z+ |G| . Let i =
1.
(ii) Find, from the corresponding reduced-order models, τ (Gji ) and ν(Gji )
for all j ∈ m, τ (|G|) and ν(|G|) as well as η z (Gij ) for all j ∈ m and
η z (|G|) for each z ∈ Z+
|G| .
(iii) Obtain Li from (6.32), ni (z) from (6.36) for each z ∈ Z+ |G| , and νi from
(6.38). Take Ni = 10 ∼ 20.
(iv) Find the largest ωni and the smallest ξi to meet (6.40), (6.42), (6.43)
and (6.44).
(v) Form hri using (6.27) and (6.35). Calculate qrii from (6.24). Let i =
i + 1. If i ≤ m, go to step (ii); otherwise end.
After qrii , i = 1, · · · , m, are determined, the ideal controller elements
IDEAL
kji can be calculated from (6.25) and (6.26). Again, they are of compli-
cated form and it is desirable to approximate them with a simpler controller
K(s) = [kij ]. Let GK = Q = [qij ] be the actual open loop transfer function
matrix. Let the frequency range of interest be
Dωi = {s ∈ C |s = jω, − ωmi < ω < ωmi } , (6.48)
6.4 Design Procedure 203
where ωmi > 0 is chosen to be large enough such that the magnitude of qrii
drops well below unity, or roughly, ωmi is well beyond the system bandwidth.
To ensure the actual loop to be close enough to the ideal one, we make
where εoi and εdi are the required performance specifications on the loop
interactions and loop performance, respectively. The values of εdi and εoi are
recommended to be in the range of 10% ∼ 30%. Smaller values of εdi and
εoi give more stringent demand on the performance and generally result in
higher order controller. One notes that although (6.50) only limits the sys-
tem couplings to a certain extent, it will still lead to almost decoupling if
the number of inputs/outputs is large. Roughly, if the number is 10, and to-
tal couplings from all other loops are limited to 30%, then each off-diagonal
element will have relative gain less than 3%, so the system is almost decou-
pled. This will be confirmed in the later simulation where almost decoupled
response is usually obtained with a εoi = 20%.
The left hand sides of (6.49) and (6.50) can be bounded respectively as
Pm m
|qii − qrii |
IDEAL
| j=1 gij (kji − kji )| X gij IDEAL
= ≤ | (kji − kji )|.
|qrii | |qrii | j=1
qrii
and
P P Pm IDEAL
j6=i |qji | j6=i | l=1 gjl (kli − kli )|
=
|qii | |qii |
Pm P IDEAL
j=1 l6=i lj (kji − kji
|g )|
≤
|qii |
ü ü
ü
m X ü
X ü g ü
= ü lj IDEAL ü
| | (kji − kji )ü .
ü qii
j=1 ü l6=i ü
One notes that (6.49) implies |qii | ≥ (1 − εdi )|qrii |, and thus, (6.49) and
(6.50) are both satisfied if for each s ∈ Dωi , there hold
X m
gij (s) IDEAL
| (kji (s) − kji (s))| ≤ di ,
j=1
qrii (s)
ü ü
m üX
X ü
ü glj (s) ü
ü | | (kji (s) − kjiIDEAL
(s))üü ≤ (1 − di )oi ,
ü qrii (s)
ü
j=1 l6=i ü
204 6. Transfer Matrix Approach
where
g (s) X glj (s)
ij
Wji (s) , max | |, | | /|qrii (s)|, (6.52)
di (1 − di )oi
l6=i
where D̃ωi is a suitably chosen subset of Dωi , and the controller elements
kji are chosen to be a rational transfer function plus a time delay in the
form of (6.45). The most important frequency range for closed-loop stability
and robustness is a decade above and below ωgi , the unity gain cross-over
frequency of the ith open loop objective transfer function qrii . Therefore,
ω
D̃ωi is chosen as the interval between ω i = 10gi and ω i = 10ωgi . In general,
the resultant controller may not be in the PID form, but this is necessary
for high performance. Note also that a possible time delay is included, and
this may look unusual. The presence of dead time in the controller could
be necessary for decoupling and it may also result from model reduction,
otherwise controller order would be even higher. Though our design needs
more calculations and a controller of high-order, only simple computations
are involved and can be made automatic. The controller designed is directly
realizable.
For actual computations, kji is found again by the model reduction al-
gorithm, with the weighting taken as W (jω) = Wji (jω) as in (6.52). When
applying the model reduction algorithm, we always set a0 = 0 in (6.45) to
impose an integrator in kji (s). Let nji be the order of the controller element
kji . To find a simplest controller, the orders of the controller elements in the
ith column, nji , j = 1, · · · , m, are determined in the following way. Firstly,
an initial estimate of nji = 2, j = 1, · · · , m, are made and then the param-
eters in kji s are calculated with the model reduction algorithm. Equations
(6.49) and (6.50) are checked. If they are satisfied, kji (s), j = 1, · · · , m are
taken as the ith column controller elements; otherwise, the following indices
are evaluated
IDEAL
ρji = max |Wji (s)(kji (s) − kji (s))|. (6.54)
s∈D̃ωi
6.4 Design Procedure 205
and let
ü
ü
∗
J = ∗ ü
j ∈ m üρj i = max ρji .
∗ (6.55)
j∈m
where estimates of τ (Gji ) and τi have already been available from Algorithm
6.4.1. Simulation shows that (6.56) usually gives a very good estimate to the
delay in kji and the search on the delay can thus be avoided.
Remark 6.4.2. It is noted that model reduction on the ideal controller is
performed over D̃ωi , which is spanned over a decade below and above the
gain cross-over frequency of qrii . The magnitude and phase of the roll-off
term ( Ni1ωni s + 1)νi in (6.27) usually remains almost unchanged over that
frequency range. Hence, although it is theoretically necessary to be included
in (6.27), it has little influence in our practical design.
The entire controller design procedure is carried out by first applying
Algorithm 6.4.1 to find the objective loop transfer function and subsequently
applying Algorithm 6.4.2 to get all the controller columns. It is recaptured
as follows for ease of reference.
(i) Apply the model reduction algorithm to |G| and Gij , i, j ∈ m, to obtain
the corresponding reduced-order models. Determine the set Z+ |G| . Let
i = 1.
(ii) Find, from the corresponding reduced-order models, τ (Gji ) and ν(Gji )
for all j ∈ m, τ (|G|) and ν(|G|), as well as η z (Gij ) for all j ∈ m and
η z (|G|) for each z ∈ Z+|G| . Obtain Li from (6.32), ni (z) from (6.36) for
+
each z ∈ Z|G| , and νi from (6.38). Take Ni = 10 ∼ 20.
(iii) Find the largest ωni and the smallest ξi to meet (6.40), (6.42), (6.43) and
(6.44). Form hri using (6.27) and (6.35). Calculate qrii from (6.24).
ω
(iv) Set the frequency range D̃ωi as [ 10gi , 10ωgi ], where ωgi is the gain cross-
IDEAL
over frequency of qrii . Evaluate kji (s), j = 1, · · · , m, from (6.25)
and (6.26). Calculate Wji (s), j = 1, · · · , m, from (6.52). Set nji = 2,
j = 1, · · · , m and J∗ = m.
(v) For each j ∈ J∗ , find nji -th order approximation, kji (s), to kji IDEAL
(s)
with the model reduction algorithm.
(vi) Check if (6.49) and (6.50) are satisfied. If yes, go to step (vii); other-
wise, calculate ρji for j ∈ m from (6.54) and find J∗ from (6.55). Set
nji = nji + 1 for each j ∈ J∗ . Go to Step (v).
(vii) Set i = i + 1, if i ≤ m, go to Step (ii), otherwise end the design, where
K(s) = [kij (s)] is a realizable controller designed for implementation.
Notice that the elements of G(s) are assumed to be strictly proper and
the elements of K(s) are designed to be proper. Thus, qji , the elements
of Q(s) = G(s)K(s) are also strictly proper, i.e., lim s→∞
Pmqji (s) = 0. Let
ωmi > 0 in (6.48) be chosen such that for all ω > ωmi , j=1 |qji (jω)| < ε
Pm
(where 0 < ε < 1), or j=1 |qji (s)| < ε for s ∈ Dωi , where Dωi denotes
Dωi ’s compliment of the Nyquist contour D, i.e., D = Dωi ∪ Dωi . We have
the following theorem.
where
ü
ü X
ü
Gier (s) = z ∈ C ü |z − qii (s)| ≤ |q (s)| .
ü ji
ü j6=i
As (6.49) and (6.50) are satisfied for s ∈ Dωi , we have for s ∈ Dωi
and
X
|qii − qrii | + |qji | ≤ di |qrii | + oi |qii | ≤ (di + oi + oi di )|qrii |.
j6=i
(6.57)
which means that the distance from qrii to the point −1 is greater that from
qrii to any point in Gier so that the Nequist curve of qrii and the ith Gersh-
gorin band Gier have the same encirclememts with respect to the point −1.
208 6. Transfer Matrix Approach
which implies that no encirclement can be made by the Gershgorin band. The
same can be said for qrii , due to the assumed |qrii (s)| < 1 for s ∈ Dωi . Thus,
the Gershgorin bands make the same number of encirclement of the point −1
as Qr = diag{qrii } does, and the system is stable as Hr = diag{hri } is so.
In the real world where the model cannot describe the process exactly,
nominal stability is not sufficient. Robust stability of the closed-loop has to
be addressed. The multivariable control system in Figure 6.1 is referred to as
being robustly stable if the closed-loop is stable for all members of a process
family. Let the actual process be denoted as G̃(s) and the nominal process
still be denoted as G(s). The following assumptions are made.
Assumption 6.5.1 The actual process Ĝ(s) and the nominal process G(s)
do not have any unstable poles, and they are both strictly proper.
Q
Consider the family of stable processes with elementwise norm-bounded
uncertainty described by
Y ü
ü ĝij (jω) − gij (jω)
ü
= Ĝ = [ĝij ] ü | | = |∆ij (jω)| ≤ γij (ω) , (6.58)
gij (jω)
i
Proof. The ith Gershgorin bands Ĝer of the actual open loop transfer matrix
Q̂(s) is,
i
Ĝer = ∪s∈D Ĝier (s),
where
ü
ü
ü X
Ĝier (s) = z ∈ C ü |z − q̂ii (s)| ≤ |q̂ji (s)|,
ü
ü j6=i
and
|z − q̂ii | ≥ |z − qrii | − |q̂ii − qrii |
Xm
= |z − qrii | − | (1 + ∆il )gil kli − qrii |
l=1
m
X
= |z − qrii | − |qii + ∆il gil kli − qrii |
l=1
m
X
≥ |z − qrii | − |qii − qrii | − | ∆il gil kli |.
l=1
Thus, we have
X X
|z − qrii | ≤ |qii − qrii | + |qji | + |∆il gil kli |.
j6=i i,j∈m
and
|z + 1| = |(1 + qrii ) + (z − qrii )| ≥ |1 + qrii | − |z − qrii |
X
≥ |1 + qrii | − (oi + di + oi di )|qrii | + |∆il gil kli | ,
i,j∈m
210 6. Transfer Matrix Approach
which will be positive, or the portion of the Gershgorin bands Gier for all
s ∈ Dωi will exclude the point −1 if
X
|1 + qrii | − (oi + di + oi di )|qrii | + |∆il gil kli | > 0,
i,j∈m
or
P
i,j∈m |∆il gil kli |
|hri (s)|(di + oi + di oi ) + < 1,
|1 + qrii |
and the above will be satisfied for all perturbation |∆ij (jω)| ≤ γij (ω), i, j ∈
m, if and only if
P
i,j∈m |γil (ω)gil (jω)kli (jω)|
|hri (jω)|(di + oi + di oi ) + < 1,
|1 + qrii (jω)|
for all −ω
Pmmi < ω < ωmi .
As j=1 |q̂ji (s)| < 1, for s ∈ Dωi and i ∈ m, thus for z ∈ Ĝier (s), where
s ∈ Dωi , we have
X
|z| − |q̂ii | ≤ |z − q̂ii | ≤ |q̂ji |,
j6=i
giving
X
|z| ≤ |q̂ii | + |q̂ji | < 1,
j6=i
or
|z + 1| > 0.
Therefore, the Gershgorin bands will exclude −1 for all s on the Nyquist
contour and will not change the number of encirclement of the point −1
compared with the nominal system. The system is robustly stable.
decoupling performance and then causing poor loop performance, too, thus
making the design fail. To control such multivariable processes, PID is not
adequate any more and higher order controller with integrator is necessary.
In this section, we will demonstrate the above observations in a great detail
with a practical example.
where ku and ωu are the ultimate gain and ultimate frequency of the process,
respectively and β is usually chosen as 4. This method is here referred to as
the SV-PID tuning.
For Multivariable PID controller design, it follows from (6.22) that g̃ii
can be regarded as the equivalent process that kii faces. kii are thus designed
with respect to the equivalent processes g̃ii with some suitable PID tuning
method, say the above SV-PID tuning. The PID parameters of the controller
off-diagonal elements are determined individually by solving (6.20) at chosen
frequencies. This method is here referred to as the MV-PID tuning.
For illustration, we will study, in details, the Alatiqi process in Example
6.1.3 with G(s) given in (6.3), and show that PID, even in its multivariable
version, is usually not adequate for complex multivariable processes for good
control performance.
4.09e−1.3s
g11 (s) = .
(33s + 1)(8.3s + 1)
212 6. Transfer Matrix Approach
−1
−2
−3
−4
−5
−3 −2 −1 0 1 2 3 4
P ID
Fig. 6.2. Nyquist Curves (—– g11 ; - - - g11 k11 )
Its Nyquist curve is shown in Figure 6.2 in solid line. Like the other elements
of the transfer matrix, g11 (s) is basically in the form of first or second order
plus time delay and simple PID controller is adequate to control it. Using the
SV-PID tuning with Am = 2 and Φm = π/4 (the default settings throughout
this chapter unless otherwise stated), a PID controller is obtained as
P ID 0.0741
k11 (s) = 2.8512 + + 9.1421s.
s
P ID
The resultant Nyquist curve of g11 (s)k11 (s) is depicted in Figure 6.2 in
dashed line and the step response of the closed-loop system is given in Figure
6.3, showing that a PID controller is sufficient for this process and there will
not be much benefits gained by using a more complex controller.
1.2
0.8
0.6
0.4
0.2
−0.2
0 50 100 150
and
g21 g12 6.93e−1.01s 6.84(33s + 1)(8.3s + 1)e−2.9s
g̃22 = g22 − = − .
g11 44.6s + 1 (31.6s + 1)(20s + 1)(45s + 1)
Their Nyquist curves are shown in Figures 6.4(a) and 6.4(b), respectively.
The SV-PID tuning gives the corresponding PID settings as
P ID 0.2890
k11 (s) = 3.1243 + + 6.7552s,
s
and
P ID 0.2825
k22 (s) = 1.0363 + + 0.7602s,
s
respectively. The resultant Nyquist curve of the loop transfer function g̃11 (s)
P ID P ID
k11 (s) and g̃22 (s)k22 (s) are shown in Figures 6.4(c) and 6.4(d), respec-
tively.
The MV-PID tuning yields
" #
P ID
3.1243 + 0.2890
s + 6.7552s −0.7896 + 0.4383
s + 0.5459s
K (s) = ,
1.7719 + 0.1723
s − 2.9869 1.0363 + 0.2825
s + 0.7602s
and the step response of the closed-loop system with this multivariable PID
controller is shown in Figure 6.5 with dashed lines. One notes that for this
2 by 2 process, the equivalent processes g̃11 (s) and g̃22 (s), though no longer
in simple first/second order plus delay form, is still in the capability of PID
control. However, one also notes that the decoupling from loop 1 to loop
2 is far from satisfactory, and this is because the Nyquist curve of a PID
controller is a vertical line in the complex plane and usually not sufficient to
compensate for the interactions.
214 6. Transfer Matrix Approach
0.6
0.4
0.4
0.2
0.2
0
0
−0.2
−0.2
−0.4 −0.4
−0.6 −0.6
−0.8 −0.8
−0.2 0 0.2 0.4 0.6 0.8 1 −0.2 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
0.4 1
0.3
0.5
0.2
0.1
0
0
−0.1
−0.5
−0.2
−0.3
−1
−0.4
−0.5 −1.5
−0.5 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4 −1.5 −1 −0.5 0 0.5 1
P ID P ID
(c) Nyquist curve of g̃11 k11 (d) Nyquist curve of g̃22 k22
1.5
y1
1
0.5
−0.5
0 50 100 150 200 250 300
1.5
y2 1
0.5
−0.5
0 50 100 150 200 250 300
Fig. 6.5. Step Response for 2 × 2 System (—– Proposed; - - - MV-PID tuning;)
To save space, we only pick up the second loop for illustration. The equivalent
process that k22 (s) faces is now
216 6. Transfer Matrix Approach
ü ü
ü 4.09e−1.3s −6.36e−0.2s −0.25e−0.4s ü
ü (33s+1)(8.3s+1) (31.6s+1)(20s+1) 21s+1 ü
ü ü
ü −4.17e−4s 6.93e −1.01s
−0.05e −5s ü
ü 34.5s+1 üü
ü 45s+1 44.6s+1
ü −1.73e−17s −11s ü
4.61e−1.02s ü
|G| ü 5.11e
13s+1 13.3s+1 18.5s+1
g̃22 (s) = 22 = ü ü ,
|G | ü 4.09e−1.3s −0.4s ü
ü (33s+1)(8.3s+1) −0.25e21s+1 ü
ü ü
ü −1.73e−17s 4.61e−1.02s üü
ü 13s+1 18.5s+1
or,
6.93e−1.01s 5.327e−15.4s
g̃22 (s) = +
44.6s + 1 (21s + 1)(45s + 1)(13.3s + 1)
0.55e−19.88s
−
(31.6s + 1)(20s + 1)(34.5s + 1)(13s + 1)
1.045e−14.98s
+
(33s + 1)(8.3s + 1)(34.5s + 1)(13.3s + 1)
122.3e−2.9s
−
(31.6s + 1)(20s + 1)(45s + 1)(18.5s + 1)
−1
18.85 0.432e−15.08s
· −
(33s + 1)(8.3s + 1)(18.5s + 1) (21s + 1)(13s + 1)
and its Nyquist curve is shown in Figure 6.6(a). The SV-PID tuning gives
P ID 0.4006
k22 (s) = 1.4929 + + 1.3907s.
s
P ID
The resultant loop Nyquist curves of g̃22 (s)k22 (s) is shown in Figure 6.6(b),
P ID
from which we find that this controller k22 (s) fails to stabilize the equivalent
process g̃22 (s). Notice that in this case, the Nyquist curves of the equivalent
process and loop transfer function are quite different from those of usual SISO
processes and they can be improper, which means that the magnitude of the
frequency response will increase and the Nyquist curve will not approach
to the origin as the frequency goes to infinity. In the proposed method, an
additional roll-off rate will be provided based on the process characteristic
such that the loop transfer function becomes proper and approximates an
objective transfer function. Obviously, these features cannot be obtained from
PID controller. More importantly, it is the complexity of the Nyquist curve of
the equivalent process over the medium frequency range that renders the PID
controller to be inadequate, as it is generally difficult to shape the Nyquist
curve of the equivalent process satisfactorily around this most important
frequency range using a PID controller.
Even if one could be satisfied with such a loop performance, PID would be
found to be yet inadequate to compensate for the interactions whose dynamics
6.6 Why PID Is Not Adequate 217
0.6 2
1.5
0.4
1
0.2
0.5
0
0
−0.2 −0.5
−1
−0.4
−1.5
−0.6
−2
−0.8
−2.5
−1 −3
−1 −0.5 0 0.5 1 1.5 2 −3 −2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2
P ID
(a) Nyquist curve of g̃22 (b) Nyquist curve of g̃22 k22
25
20 −1
−2
15
−3
10 −4
−5
5
−6
−7
0
−8
−5 −9
−4 −3 −2 −1 0 1 2 3 4 5 −0.5 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2
IDEAL
(c) Nyquist curves (—– k32 ;-- (d) Nyquist curves (—– qr22 ; - - -
P ID HIGH
- k32 ) g̃22 k22 )
may also be very complicated and it would fail to achieve the decoupling. For
an illustration, according to (6.20), the controller off-diagonal element, say
k32 (s), should be
ü ü
ü 4.09e−1.3s −6.36e−0.2s ü
ü (33s+1)(8.3s+1) (31.6s+1)(20s+1) ü
ü ü
ü −1.73e−17s 5.11e −11s ü
ü 13s+1 13.3s+1
ü
IDEAL
k32 (s) = − üü −1.3s
ü k22 (s),
−0.4s ü
(6.59)
4.09e −0.25e
ü (33s+1)(8.3s+1) 21s+1 ü
ü ü
ü −1.73e−17s 4.61e−1.02s üü
ü 13s+1 18.5s+1
P ID
with k22 (s) = k22 (s). The MV-PID design method gives the PID form of
P ID
the off-diagonal controller elements k32 (s) as
218 6. Transfer Matrix Approach
P ID 0.2152
k32 (s) = −1.6825 − − 2.381s.
s
IDEAL
The Nyquist curve of the ideal decoupler k32 (s) and its PID approxima-
P ID
tion k32 are shown in Figure 6.6(c) in solid line and dash line, respectively,
from which one sees that the Nyquist curve of the ideal decoupler k32 (s)
P ID
is quite complicated and thus its approximation by k32 (s) of PID form is
inadequate. Actually, with the MV-PID tuned controller:
5.976 + 0.76 0.61
s + 11.7s 2.16 + s + 2.72s −0.825 +
0.218
s + 0.994s
K P ID (s) = 0.456 0.41 0.133
2.59 + s − 5.87s 1.49 + s + 1.39s 0.56 + s + 0.328s ,
1.5
y1
1
0.5
0
0 100 200 300 400 500 600
1.5
y2
1
0.5
0
−0.5
0 100 200 300 400 500 600
2
y3
1
−1
0 100 200 300 400 500 600
equivalent processes. Of course, One can always de-tune the PID controllers,
or reduce the gains to a sufficient degree to generate a stable closed-loop if
the original process is stable. However, we have to have the sluggish response
and poor performance, which might not be desirable or acceptable.
With our method, an objective loop transfer function for the second loop
for this example is specified as
0.17902 e−6.62s
qr22 (s) = ,
s2 + 2 × 0.1790 × 0.707s + 0.17902 (1 − e−6.62s )
HIGH
and a 6th-order controller, k22 (s), is designed as
HIGH
such that g̃22 (s)k22 (s) approximates qr22 (s). The Nyquist curve of the
IDEAL
actual loop transfer function g̃22 (s)k22 (s) and the ideal one qr22 (s) are
shown in Figure 6.6(d) with dashed line and solid line, respectively. We can
HIGH HIGH
see that g̃22 (s)k22 (s) with the 6th-order k22 (s) yields very close fitting
to the ideal loop transfer function in frequency domain while generally PID
controller is not capable of that.
To achieve decoupling, according to (6.20), the controller off-diagonal el-
HIGH
ement, say k32 (s), should be given by (6.59) with k22 (s) = k22 (s). The
IDEAL
Nyquist curve of the ideal decoupler k32 (s) is shown in Figure 6.8(a)
with solid line, which is again a function that PID is difficult to approximate.
HIGH
A 6th-order approximation k32 (s) is obtained as
and its Nyquist curve is shown in Figure 6.8(a) with dashed line, from which
we see that it approximates the ideal one very well over the interested fre-
quency range. The other controller elements are obtained as
HIGH −4433s4 − 1021s3 + 679.6s2 + 50.6s + 1
k11 (s) = ,
10580s4 + 2104s3 + 1006s2 + 6.468s
The time domain step response of the overall multivariable feedback system
is shown in Figure 6.7 with solid lines, from which we see that the general
controller gives very good loop and decoupling performance.
0.1
5
0
4
−0.1
3
−0.2
2
−0.3
1
−0.4
0
−0.5
−1 −0.6
−2
−2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5 0 0.5 1 1.5
IDEAL
(a) Nyquist curves (—– k32 ;- (b) Nyquist curve of g̃22
HIGH
- - k32 )
2 20
1.5
0
1
0.5 −20
0
−40
−0.5
−60
−1
−1.5
−80
−2
−100
−2.5
−3 −120
−4 −3 −2 −1 0 1 2 3 4 −70 −60 −50 −40 −30 −20 −10 0 10 20
P ID IDEAL
(c) Nyquist curve of g̃22 k22 (d) Nyquist curves (—– k32 ; -
P ID
- - k32 )
Its Nyquist curve is shown in Figure 6.8(b), and the complexity of the equiv-
alent process is obvious. The SV-PID tuning gives the PID controller as
P ID 0.7503
k22 (s) = 3.5264 + + 4.1438s.
s
P ID1
The resultant loop Nyquist curves of g̃22 (s)k22 (s) is shown in Figure 6.8(c),
from which we can expect very poor loop performance or instability.
Apart from the poor loop performance, PID is also found to be inadequate
to achieve decoupling for this process. To achieve decoupling, (6.20) applies,
say for k32 (s), to get
ü ü
ü 4.09e−1.3s −6.36e−0.2s −0.49e−5s ü
ü (33s+1)(8.3s+1) (31.6s+1)(20s+1) 22s+1 ü
ü ü
ü −1.73e−17s 5.11e −11s
−5.48e −0.5s ü
ü 13s+1 13.3s+1 15s+1 ü
ü ü
ü −11.18e−2.6s −0.02s −0.6s ü
ü (43s+1)(6.5s+1) (45s+1)(10s+1) (48s+1)(6.3s+1) ü
14.04e 4.49e
IDEAL
k32 (s) = − ü ü k22 (s),
ü 4.09e−1.3s −0.25e−0.4s −0.49e−5s ü
ü (33s+1)(8.3s+1) 21s+1 22s+1 ü
ü ü
ü −1.73e−17s 4.61e−1.02s −5.48e−0.5s ü
ü 13s+1 18.5s+1 15s+1 ü
ü ü
ü −11.18e−2.6s −0.05s −0.6s ü
ü (43s+1)(6.5s+1) (31.6s+1)(5s+1)
−0.1e 4.49e
(48s+1)(6.3s+1)
ü
(6.60)
P ID
with k22 (s) = k22 (s). The MV-PID design method gives the PID form of
P ID
the off-diagonal controller elements k32 (s) as
P ID 1.9152
k32 (s) = −3.7326 + − 3.5149s.
s
IDEAL
The Nyquist curve of the ideal decoupler k32 (s) and its PID approxima-
P ID
tion k32 are shown in Figure 6.8(d) in solid line and dash line, respectively,
222 6. Transfer Matrix Approach
HIGH
is designed such that g̃22 (s)k22 (s) approximates qr22 (s). The Nyquist
HIGH
curve of the actual loop transfer function g̃22 (s)k22 (s) and the ideal loop
one qr22 (s) are shown in Figure 6.9(a) with dash line and solid line, respec-
HIGH
tively, and they are different. Note that k22 (s) contains a time delay term.
This can happen and is necessary in the multivariable case for decoupling,
as shown before. If this delay is removed or reduced, or in other word, the
time delay in the second decoupled loop transfer function qr22 is reduced,
then one or some controller off-diagonal element will have time predictions
(eLs , L > 0), which is not realizable.
1 1
0
0
−1
−1
−2
−2
−3
−3 −4
−4 −5
−6
−5
−7
−6
−0.5 −0.4 −0.3 −0.2 −0.1 0 0.1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4
IDEAL
(a) Nyquist curves (—– qr22 ; - - - (b) Nyquist curves (—– k32 ;-
HIGH HIGH
g̃22 k22 ) - - k32 )
with solid line, which is again a function that PID is difficult to approximate.
HIGH
A 5th-order approximation k32 (s) is obtained with the proposed method
as
66110s5 − 841.8s4 − 107.1s3 + 216.1s2 + s − 0.1403 −1.683s
k32 (s) = e ,
−619800s5 − 102700s4 − 10680s3 − 545.2s2 − 14.23s
and its Nyquist curve is shown in Figure 6.9(b) with dash line, from which we
see that it approximates the ideal one very well over the interested frequency
range. The remaining controller elements generated by the proposed method
are listed in (6.61) to (6.74). The resultant time domain step response is
shown in Figure 6.10, from which we see that the general controller gives
excellent loop and decoupling performance.
y1
1
0.5
0
−0.5
0 200 400 600 800 1000 1200 1400 1600 1800 2000
y2 1
0.5
0
−0.5
0 200 400 600 800 1000 1200 1400 1600 1800 2000
y3 1
0.5
0
−0.5
0 200 400 600 800 1000 1200 1400 1600 1800 2000
y4 1
0.5
0
−0.5
0 200 400 600 800 1000 1200 1400 1600 1800 2000
Fig. 6.10. Step response for Example 6.1.3
very difficult to compensate for when the I/O number of the processes
increases. The resultant equivalent processes can be extremely compli-
cated even when all the individual channel dynamics are of simple first
or second order plus time delay form.
(ii) For such complex multivariable processes, PID controller is usually in-
adequate, not only because the equivalent processes is difficult to be
controlled by PID, but also the interactions are too difficult to be com-
pensated for by PID.
(iii) Existing non-PID control are usually state-space based. Their specifi-
cations are generally quite different from those used in process control.
Time delay is frequently encountered in process control while state-space
based approach usually cannot handle the time delay easily. The decou-
pling requirement which is very important for process control is usually
not considered explicitly in state-space based approaches. These explain
for discrepancy between the many well developed state-space based mul-
tivariable designs (H∞ , LQG, ...) and the rare use of them in process
control.
(iv) A general, effective and efficient design of non-PID control for com-
plex multivariable process capable of achieving high performance, is thus
needed and should have great potential for industrial applications. This
chapter has presented such a method.
230.5s + 17.25
G22 (s) = e−5.847s ,
1718s2 + 92.27s + 1
817.3s2 − 4474s − 20.48
G23 (s) = e−14.81s ,
504200s3 + 22520s2 + 288.9s + 1
12630s2 + 376.3s + 14.45
G31 (s) = e−8.905s ,
68680s3 + 8196s2 + 420.1s + 1
157.6s + 4.849
G32 (s) = e−4.334s ,
2069s2 + 97.88s + 1
−348s2 + 266.3s + 0.5485
G33 (s) = e−0.8209s .
121800s3 + 28450s2 + 478.9s + 1
We got Z+ |G| = {3.027} from |G(s)|. Set i = 1.
(ii) From the above reduced order models, we obtain τ (|G|) = 15.3,
τ (G11 ) = 0.946, τ (G12 ) = 11.47, τ (G13 ) = 14.77; for z = 3.027 in Z+
|G| ,
ηz (|G|) = 1 and ηz (G1j ) = 0, j = 1, 2, 3; τ1 = 0.9463 and L1 = 14.35.
And n1 (z) = 1.
(iii) Take ξ1 = 0.707. With φm1 = π/4, ωg1 is obtained from (6.43) as
ωg1 = 0.0529, and ωn1 is determined from (6.42) as ωn1 = 0.0794. hr1
is then formed from (6.27) as
−689.7s2 + 57.31s + 1
k11 (s) = ,
710.8s2 + 64.14s
20330s2 + 235.2s + 1 −10.52s
k21 (s) = e ,
319900s2 + 1400s
−50380s2 + 462.1s + 1 −13.82s
k31 (s) = e .
−4709000s2 − 74420s
(vi) It is found that (6.49) and (6.50) are not satisfied. ρ11 , ρ21 and ρ31 are
then calculated as 6.965, 6.858 and 4.400 from (6.54) and J∗ is thus
found as {1} from (6.55). Set n11 = 3, i.e., the order of k̂11 (s) increases
by 1. Go back to step (v). This procedure repeats until (6.49) and (6.50)
are both satisfied and the following k11 , k21 and k31 are obtained:
226 6. Transfer Matrix Approach
For i = 3, we have
−1430s4 + 938.5s3 + 29.26s2 + s + 0.008427 −8.084s
k13 (s) = e ,
34560s4 + 2527s3 + 485.3s2 + 1.039s
−49.8s3 + 6.178s2 + s + 0.02832 −3.513s
k23 (s) = e ,
554.6s3 + 62.99s2 + 11.34s
−32.71s3 − 3.641s2 + s + 0.03188
k33 (s) = .
4964s3 + 411s2 + 102.3s
If we restrict the controller to the PID type to approximate the ideal
controller, the design procedure produces
P ID
K =
0.0152 0.0475 0.0022
−0.0442 + s + 1.2413s −0.0095 − s − 0.7917s −0.0445 + s + 0.2294s
0.0062 + 0.0006
+ 0.0715s 0.0998 + 0.0107
− 0.4601s 0.0198 + 0.0021
− 0.2532s
s s s .
0.0356 0.0113 0.0005
−0.0000538 + 1000s + 0.0188s 0.0342 − s − 0.2112s −0.0073 + s + 0.0298s.
The step response of the closed-loop system is shown in Figure 6.11, where
the solid line is the response with K(s) and the dash line with K P ID (s), from
which we can see that the performance of the general controller is very satis-
factory while the PID performance is not. The response for the manipulated
variables are shown in Figure 6.12, from which we see that the manipulated
variables are quite small and smooth for the high-order controller while PID
control generates big peaks from the derivative terms.
In order to see the robustness of the proposed method, we increase the
static gains in all the elements in the process transfer matrix by 40%, i.e.,
each gij (s) is perturbed to 1.4gij (s). The step response under such gain per-
turbations is shown in Figure 6.13 with dashdot lines, while the nominal
6.7 More Simulation 227
1.5
y1
1
0.5
0
0 100 200 300 400 500 600 700 800 900 1000
y2 1
0.5
0
0 100 200 300 400 500 600 700 800 900 1000
y3
1
0.5
−0.5
0 100 200 300 400 500 600 700 800 900 1000
Fig. 6.11. Output Step Responses for Example 6.1.1 (—– High order; - - - PID)
u1 1
0
−1
−2
0 100 200 300 400 500 600 700 800 900 1000
0.6
u2
0.4
0.2
0
−0.2
0 100 200 300 400 500 600 700 800 900 1000
u3 0
−0.2
−0.4
0 100 200 300 400 500 600 700 800 900 1000
Fig. 6.12. Input Step Responses for Example 6.1.1C (—– High order; - - - PID)
228 6. Transfer Matrix Approach
y1
1
0.5
0
0 100 200 300 400 500 600 700 800 900 1000
y2
1
0.5
0
0 100 200 300 400 500 600 700 800 900 1000
y3
1
0.5
−0.5
0 100 200 300 400 500 600 700 800 900 1000
performance is shown in solid lines. One notes that this gain variation does
not affect the decoupling, which is expected. To introduce variation in the
system dynamics, we perturb the nominal process by increasing the time
constants in all gij (s) by 40%. The corresponding step response is shown in
Figure 6.13 with dashed lines, exhibiting the deterioration of both the loop
and decoupling performance. However, such deterioration is reasonable as
compared with the “size” of the perturbation.
Example 6.1.2 (continued). Consider the 4 by 4 Doukas process in
(6.2). Following our method, the objective closed-loop transfer functions are
determined as
0.044422 e−17.78s
hr1 (s) = ,
s2 + 2 × 0.04442 × 0.707s + 0.044422
0.045272 e−17.45s
hr2 (s) = ,
s2 + 2 × 0.04527 × 0.707s + 0.045272
0.045192 e−17.48s
hr3 (s) = ,
s2 + 2 × 0.04519 × 0.707s + 0.045192
0.096682 e−8.17s
hr4 (s) = .
s2 + 2 × 0.09668 × 0.707s + 0.096682
6.7 More Simulation 229
y1 1
0.5
0
−0.5
0 200 400 600 800 1000 1200
y2 1
0.5
0
−0.5
0 200 400 600 800 1000 1200
y3 1
0.5
0
−0.5
0 200 400 600 800 1000 1200
y4 1
0.5
0
−0.5
0 200 400 600 800 1000 1200
Fig. 6.14. Step Responses for Example 6.1.2C
−12.21s2 + s + 0.02426
k13 (s) = , (6.63)
1982s2 + 50.13s
0.20572 e−5.760s
hr1 (s) = ,
s2 + 2 × 0.2057 × 0.707s + 0.20572
0.057162 e−20.73s
hr2 (s) = ,
s2 + 2 × 0.05716 × 0.707s + 0.057162
232 6. Transfer Matrix Approach
0.13532 e−8.760s
hr3 (s) = ;
s2 + 2 × 0.1353 × 0.707s + 0.13532
and
−0.6015s2 + s + 0.09952 −2s
k11 (s) = e ,
−2.783s2 − 1.409s
3.685s3 + 4.08s2 + s + 0.08144 −1.861s
k12 (s) = e ,
5.537s3 + 0.793s2 + 0.1084s
−0.1618s3 + 1.721s2 + s + 0.1044 −2s
k13 (s) = e ,
2.406s3 + 0.8755s2 + 0.2685s
−0.0771s3 + 0.5865s2 + s + 0.1633 −3.513s
k21 (s) = e ,
−0.09085s3 − 0.06274s2 − 0.02637s
−0.5531s3 + 2.607s2 + s + 0.1366 −3.494s
k22 (s) = e ,
−0.09555s3 − 0.01373s2 − 0.00194s
4.332s3 + 2.386s2 + s − 0.1834 −3.936s
k23 (s) = e ,
0.1233s3 + 0.04899s2 + 0.01975s
−0.5639s2 + s + 0.05756
k31 (s) = ,
−1.875s2 − 1.051s
−4.014s4 + 7.842s3 + 4.624s2 + s + 0.0617
k32 (s) = ,
−7.849s4 − 5.138s3 − 0.7457s2 − 0.08274s
−0.306s3 + 2.29s2 + s + 0.07093
k33 (s) = .
2.016s3 + 0.6972s2 + 0.2153s
The step response is shown in Figure 6.15.
Example 6.1.5 (continued). Consider the depropanizer process in
(6.5). It follows from the design procedure that
and
−53.59s3 + 24.86s2 + s + 0.005836
k11 (s) = ,
4217s3 + 85.86s2 + 1.091s
6.7 More Simulation 233
y1
1
0.5
−0.5
0 100 200 300 400 500 600
y2
1
0.5
−0.5
0 100 200 300 400 500 600
y3
1
0.5
−0.5
0 100 200 300 400 500 600
y1
1
0.5
−0.5
0 1000 2000 3000 4000 5000 6000
y2
1
0.5
−0.5
0 1000 2000 3000 4000 5000 6000
y3
1
0.5
−0.5
0 1000 2000 3000 4000 5000 6000
6.8 Conclusions
For plants with time delay, frequency domain methodologies are popular,
such as the inverse Nyquist array (Rosenbrock, 1974) and characteristic lo-
cus method (Macfarlane, 1980). Wang et al. (1997a) derived the decoupling
conditions for delay systems. The approach presented in this chapter is based
on Zhang (1999). In many cases, one replaces time delay by a rational func-
tion, say via Pade approximation, and then adopts delay-free designs (Morari
and Zafiriou, 1989). Such approximations will deteriorate when time delay
increases, and time delay compensation will be necessary to get high perfor-
mance and a topic of the next chapter.
7. Delay Systems
where
and gij0 (s) are strictly proper, stable, and scalar rational functions and Lij
are non-negative constants. It is also assumed that a proper input-output
pairing has been made for G(s) such that none of the m principal minors of
G(s) (the i th principle minor is the determinant of a (m−1)×(m−1) matrix
obtained by deleting the i th row and i th column of G(s)) is zero. Our task in
this section is to characterize all the realizable controllers K and the resultant
closed-loop transfer functions such that the IMC system is internally stable
and decoupled.
The closed-loop transfer matrix H between y and r can be derived from
Figure 7.1 as
H = GK[I + (G − Ĝ)K]−1 ,
which becomes
H = GK, if Ĝ = G.
N (s)
G(s) = ,
d(s)
adj N (s)
K(s) = ,
σ(s)
where adj N (s) is the adjoint matrix of N (s) and σ(s) is a stable polynomial
and its degree is high enough to make the rational part of each kij proper.
Then, this K(s) is stable and realizable and it decouples G(s) since
|N (s)|
G(s)K(s) = Im ,
d(s)σ(s)
Theorem 7.1.1. For a stable, square and multi-delay process G(s), the de-
coupling problem with stability via the IMC is solvable if and only if G(s) is
non-singular.
where
|G|
g̃ii = , ∀ i ∈ m; (7.3)
Gii
and
where
Gij
ψji = , ∀ i, j ∈ m, j = i. (7.5)
Gii
240 7. Delay Systems
One notes that for a given i, the resulting diagonal element of GK is indepen-
dent of the controller off-diagonal elements but contains only the controller
diagonal element kii .
To demonstrate how to find g̃ii and ψji , take the following as an example:
1 e−2s −1 e−6s
s + 2 s +2
G(s) = 2 . (7.6)
s − 0.5 e−3s (s − 0.5) e−8s
2 3
(s + 2) 2(s + 2)
Simple calculations give
2(s − 0.5)(s + 2)e−9s + (s − 0.5)2 e−10s
|G| = ,
2(s + 2)4
e−2s (s − 0.5)e−3s
G22 = , G12 = − .
s+2 (s + 2)2
It follows from (7.3) that the decoupled loops have their equivalent processes
as
2(s + 2)e−s + (s − 0.5)e−2s
g̃11 = ,
(s − 0.5)(s + 2)
7.1.1 Analysis
In what follows, we will develop the characterizations of time delays and non-
minimum phase zeros on the decoupling controller and the resulting decou-
pled loops. For a nonsingular delay process G(s) in (7.1), a general expression
for g̃ii in (7.3) will be
M −αk s
k=0 nk (s)e
φ(s) = N , (7.7)
d0 (s) + l=1 dl (s)e−βl s
where nk (s) and dl (s) are all non-zero scalar polynomials of s, α0 < α1 <
· · · < αM and 0 < β1 < β2 < · · · < βN . Define the time delay for non-zero
φ(s) in (7.7) as
τ (φ(s)) = α0 .
It is easy to verify τ (φ1 φ2 ) = τ (φ1 ) + τ (φ2 ) and τ (φ−1 (s)) = −τ (φ(s)) for
any non-zero φ1 (s), φ2 (s) and φ(s). If τ (φ(s)) ≥ 0, it measures the time
required for φ(s) to have a non-zero output in response to a step input. If
τ (φ(s)) < 0 then the system output will depend on the future values of input.
It is obvious that for any realizable and non-zero φ(s), τ (φ(s)) can not be
negative. Therefore, a realizable K requires
τ (kji ) ≥ 0, ∀ i ∈ m, j ∈ Ji , (7.8)
which, by
|G|
τ (g̃ii ) = τ = τ (|G|) − τ (Gii ),
Gii
becomes
It follows from (7.9) that k11 and k22 are characterized on their time delays
by
τ (k11 ) ≥ τ (G11 ) − τ1 = 8 − 3 = 5,
τ (k22 ) ≥ τ (G22 ) − τ2 = 2 − 2 = 0.
τ (h11 ) ≥ τ (|G|) − τ1 = 9 − 3 = 6,
τ (h22 ) ≥ τ (|G|) − τ2 = 9 − 2 = 7.
ηz (φ−1 ) = −ηz (φ) for any non-zero transfer function φ1 (s), φ2 (s) and φ(s).
Obviously, a non-zero transfer function φ(s) is stable if and only if ηz (φ) ≥ 0,
∀ z ∈ C+ .
A stable K thus requires
ηz (kji ) ≥ 0, ∀ i ∈ m, j ∈ Ji , z ∈ C+ . (7.12)
or
where
Since G(s) is stable, so are Gij . One sees that for ∀ i ∈ m and ∀ z ∈ C+ ,
ηi (z) ≥ 0. We have, for ∀ i ∈ m and ∀z ∈ C+ ,
and
Equation (7.15) also implies that kii need not have any non-minimum
phase zeros except at z ∈ Z+
Gii . Therefore, by (7.13), kii is characterized on
its non-minimum phase zeros by
which, by
|G|
ηz (g̃ii ) = ηz = ηz (|G|) − ηz (Gii ),
Gii
becomes
and
m
|G| 1
ηz (|G|) − ηi (z) = ηz = ηz gij Gij
(s − z)ηi (z) (s − z)ηi (z) j=1
m ij
G
= ηz gij ≥ 0.
j=1
(s − z)ηi (z)
i.e., the second loop must contain a non-minimum phase zero at z = 0.5 of
multiplicity one while the first loop need not contain any non-minimum phase
zero.
Theorem 7.1.2. If the IMC system in Figure 7.1 is decoupled, stable and
realizable, then (i) the controller K’s diagonal elements are characterized by
(7.9) on its time delays and by (7.16) on its RHP zeros, and they uniquely
determine the controller’s off-diagonal elements by (6.20); (ii) the diagonal
elements of GK, hii = g̃ii kii , are characterized by (7.11) on its time delays
and (7.19) on its RHP zeros.
1
k11 = k22 = ,
s+ρ
which violates the conditions of Theorem 7.1.2. It gives rise to
−2(s + 2)
k21 = e5s ,
(s + ρ)(s − 0.5)
which is neither realizable nor stable. The resultant h11 is given by
2s + 4 + (s − 0.5)e−s −s
h11 = g̃11 k11 = e ,
(s + ρ)(s + 2)(s − 0.5)
which is unstable, too. Such kii should be discarded.
A quite special phenomenon in such a decoupling control is that there
might be unstable pole-zero cancellations in forming g̃ii kii . In contrast to
the normal intuition, this will not cause instability. Notice from the previous
example that g̃11 has a pole at s = 0.5 and k11 has at least one zero at the
same location as ηz (k11 )|z=0.5 ≥ 1. Hence, an unstable pole-zero cancellation
at z = 0.5 occurs in forming g̃11 k11 . However, the resulting closed-loop system
is still stable because all the controller diagonal and off-diagonal elements are
stable. In fact, as far as G and K are concerned, there is no unstable pole-zero
cancellation since both G and K are stable.
7.1.2 Design
In this subsection, practical design of IMC is considered. We first briefly
review the IMC design for SISO case. Let g(s) be the transfer function
model of a stable SISO process and k(s) the controller. The resulting closed-
loop transfer function in case of no process-model mismatch is h(s), where
h(s) = g(s)k(s). It is quite clear that if h(s) has been specified then k(s)
is determined. However h(s) can not be specified arbitrarily. In IMC de-
sign, the process model g(s) is first factored into g(s) = g + (s)g − (s), where
g + (s) contains all the time delay and RHP zeros. Then h(s) is chosen as
h(s) = g + (s)f (s), where f (s) is the IMC filter, and consequently k(s) is
given by k(s) = h(s)g −1 (s).
The factorizations which yield g + (s) are not unique. Holt and Morari
(1985a,b) suggested that the following form be advantageous:
n
zi − s
g + (s) = e−Ls , (7.20)
i=1
zi + s
where L is the time delay and z1 , z2 , · · · , zn are all the RHP zeros present
in g(s). The IMC filter f (s) is usually chosen in the form:
1
f (s) = ,
(τ s + 1)l
where τ and l are the filter time constant and order, respectively. The filter
time constant τ is a tuning parameter for performance and robustness.
7.1 The IMC Scheme 247
z−s
ηz (|G|)−ηi (z)
hii = e−(τ (|G|)−τi )s fi (s), ∀ i ∈ m, (7.21)
z+s
z∈Z+
|G|
where fi (s) is the i-th loop IMC filter. This choice of hii (s) will then determine
kii (s) as
−1
kii (s) = hii (s)g̃ii (s), (7.22)
Remark 7.1.1. Good control performance of the loops is what we seek apart
from the loop-decoupling requirement. However, we do not explicitly define
a performance measure and design the controller via the minimization (or
maximization) of some measure. These are done in model predictive control,
where the “size” of the error between the predicted output and a desired
trajectory is minimized; and in LQG control, where a linear quadratic index
is minimized to yield an optimal controller. These approaches have a dif-
ferent framework than ours. What we do here is to follow the IMC design
approach with the requirement of decoupling for the control of multivariable
multi-delay processes. Loop performance is simply observed with closed-loop
step responses for each output, in terms of traditional performance specifica-
tions such as overshoot, rise time and settling time in comparison with the
existing schemes of similar purposes. A link of the IMC design theory with
the quantitative error measure is that the item with the non-minimum phase
factor g + (s) in (7.20) used in IMC closed-loop transfer function in (7.21) can
actually minimize the H-2 norm of the error signal in the presence of a step
reference change (Morari and Zafiriou, 1989).
Remark 7.1.2. One may note that there is the case where good decoupling
and performance may not be both achievable and one is reached at cost of an-
other. For example, decoupling should never be used in flight control, instead,
couplings are deliberately employed to boost performance. This case is how-
ever not the topic of the present paper. We here address applications where
decoupling is required. Indeed, decoupling is usually required (Kong, 1995)
248 7. Delay Systems
and the controller K are stable, as discussed before. In the real world where
the model does not represent the process exactly, nominal stability is not
sufficient and robust stability of the IMC system has to be ensured. As a
standard assumption in robust analysis, we assume that the nominal IMC
system is stable (i.e., both Ĝ and K are stable). Suppose that the actual
process G is also stable and is described by
G∈ = {G : σ [G(jω) − Ĝ(jω)]Ĝ−1 (jω) = σ (∆m (jω)) ≤ ˜lm (ω)},
where ˆlm (ω) is the bound on the multiplicative uncertainty. It follows (Chap-
ter 3) that the IMC system is robust stable if and only if
−1
σ(Ĝ(jω)K(jω)) < lm (jω), ∀ω. (7.23)
Let K ∗ be the ideal controller obtained from (7.22) and (7.4), we have ĜK ∗ =
diag {hii }. Thus (7.23) becomes
σ diag {hii (jω)}K ∗ −1 (jω)K(jω) < lm −1
(jω), ∀ω.
One notes from (7.21) that diag {hii (jω)} = U diag {fi (jω)}, where U is a
unitary matrix. Noticing σ(U A) = σ(A) for a unitary U , thus a sufficient
and necessary robust stability condition is obtained as
σ diag {fi (jω)}K ∗ −1 (jω)K(jω) < lm −1
(jω), ∀ω. (7.24)
In practice, we can evaluate and plot the left hand side of (7.24) and compare
−1
with lm (jω) to see whether robust stability is satisfied given the multiplica-
tive uncertainty bound lm (jω). One also notes that if the error between the
ideal controller K ∗ and its reduced-order model K are small, i.e., K ∗ ≈ K,
then the robust stability condition becomes
−1
σ (diag {fi (jω)}) < lm (jω), ∀ω.
which indicates that the IMC filter should be chosen such that its largest
singular value is smaller than the inverse of the uncertainty magnitude for
all ω.
7.1.3 Simulation
In this subsection, several simulation examples are given to show the ef-
fectiveness of the decoupling IMC design. The effects of time delays and
non-minimum phase zeros are illustrated and control performance as well as
its robustness is compared with the existing multivariable Smith predictor
schemes. We intend to use commonly cited literature examples with real live
250 7. Delay Systems
1.5
y1
0.5
0
0 10 20 30 40 50 60 70 80 90 100
1.5
y2
1
0.5
−0.5
0 10 20 30 40 50 60 70 80 90 100
Step (i ) The application of model reduction to |G|, G11 and G12 produces
−0.1077s2 − 4.239s − 0.3881 −6.3s
|Ĝ| = e ,
s2 + 0.1031s + 0.0031
−1.349s − 42.81 −3s
Ĝ11 = e ,
s2 + 31.85s + 2.207
−0.6076s − 17.66 −7s
Ĝ12 = e .
s2 + 29.26s + 2.676
It is clear that τ (|Ĝ|) = 6.3, τ (Ĝ11 ) = 3, and τ (Ĝ12 ) = 7. Thus, by (7.10),
one finds τ1 = min{3, 7} = 3. Since |Ĝ| is of minimum phase, Z+ |Ĝ|
= ∅ and
there is no need to calculate η1 (z).
−3.3s
Step (ii ) According to (7.21), we have h11 = e s+1 with the filter chosen
1 −1
as s+1 . k11 is then obtained as k11 = g̃11 h11 from (7.22). The application of
model reduction to k11 yields
0.3168s2 + 0.0351s + 0.0012 −0.9s
k̂11 = e .
s2 + 0.1678s + 0.0074
Step (iii ) k21 is calculated from (7.4). Applying model reduction to k21
yields
0.1405s2 + 0.0180s + 0.0007 −4.9s
k̂21 = e .
s2 + 0.2013s + 0.0123
Step (iv ) The smallest time delay in k̂11 and k̂21 is 0.9 and then subtracted
from both elements’ time delays to generate the final k̂11 and k̂21 .
−5.3s
Step (v ) Loop two is treated similarly, where h22 = e s+1 . This results in
the overall controller as
0.3168s2 + 0.0351s + 0.0021 −0.2130s2 − 0.0222s − 0.0008 e−2s
2
s + 0.1678s + 0.0074 2
s + 0.1399s + 0.0051
K=
2 2
.
0.1405s + 0.0180s + 0.0007 e−4s −0.1798s − 0.0206s − 0.0008
2 2
s + 0.2031s + 0.0123 s + 0.1607s + 0.0073
1
10
0
10
−1
10
−1 0
10 10
w
y1 2
1.5
0.5
0
0 10 20 30 40 50 60 70 80 90 100
y2 2
1.5
0.5
−0.5
0 10 20 30 40 50 60 70 80 90 100
In order to see the robustness of the IMC design, we increase the dead
times of the process’ diagonal elements by 15%. In order to check the stability
of the perturbed system, the left hand side and right hand side of (7.24) are
plotted in Fig 7.3 with solid line and dashed line respectively. From the figure,
we conclude that the IMC system remains stable under the given perturba-
tions. The set-point responses are shown in Figure 7.4, exhibiting that the
performance of the IMC design is much superior compared with the Jerome’s
multivariable Smith predictor scheme. The resulting ISE is calculated as 7.76
for the IMC method while it is infinite for Jerome’s scheme due to instability.
♦
where the delay free parts gij0 (s) are identical to Example 7.1.1. It follows
from the IMC procedure that
One readily sees that τ (|Ĝ|) = 13.6, τ (Ĝ11 ) = 15, τ (Ĝ12 ) = 2 and thus
τ1 = min{15, 2} = 2 from (7.10). A difference from Example 7.1.1 is
that |Ĝ| has a RHP zero at z = 0.136 with a multiplicity of one, i.e.,
Z+|Ĝ|
= {0.136} and ηz (|Ĝ|)|z=0.136 = 1. It is easy to find ηz (Ĝ11 )|z=0.136 = 0
and ηz (Ĝ12 )|z=0.136 = 0 since Ĝ11 and Ĝ12 are both of minimum phase.
We thus have η1 (z)|z=0.136 = min{0, 0} = 0 from (7.14) and h11 =
0.136−s −11.6s 1
(0.136+s)(s+1) e from (7.21) with the filter of s+1 . For the second loop,
1 0.136−s
with the filter chosen as s+1 , h22 is obtained as h22 = (0.136+s)(s+1) e−12.6s .
The rest of design is straightforward and gives
0.3656s2 + 0.0524s + 0.002 e−13.1s −0.2375s2 − 0.0294s − 0.0011 e−8s
s + 0.2265s + 0.0124
2 2
s + 0.1790s + 0.0070
K=
2 2
.
0.1551s + 0.0291s + 0.012 −0.2003s − 0.0317s − 0.0014
s2 + 0.2726s + 0.0226 s2 + 0.2215s + 0.0136
y1
1.5
0.5
−0.5
−1
0 50 100 150 200 250 300
y2 1.5
0.5
−0.5
−1
0 50 100 150 200 250 300
its response is very close to that of the ideal IMC controller (dotted line). The
resulting ISE is calculated as 52.5, 71.16 and 49.4 for the IMC, Ogunnaike
and Ray’s scheme, and the ideal IMC, respectively. ♦
Example 7.1.3. The process studied by Luyben (1986a) has the following
transfer function matrix:
0.126e−6s −0.101e−12s
60s + 1
G(s) = (45s + 1)(48s + 1) .
0.094e−8s −0.12e−8s
38s + 1 35s + 1
It follows from the design procedure that with filters chosen as f1 (s) =
−6.5s −8.5s
1
f2 (s) = 3s+1 , h11 and h22 are obtained as h11 = e3s+1 and h22 = e3s+1 ,
respectively. The resultant controller is
146.8s2 + 6.355s + 0.0733 −6.027s2 − 0.3416s − 0.0053 e−9.3s
2 2
K = s +20.3465s + 0.0034 s + 0.0395s + 0.0003
.
104.5s + 4.719s + 0.0561 −68.2s2 − 3.490s − 0.0527
s2 + 0.34s + 0.0034 s2 + 0.2496s + 0.0024
Since the multivariable Smith predictor design for this process is not avail-
able in the literature, the BLT tuning method (Luyben, 1986a) is used for
7.1 The IMC Scheme 255
1.5
y1
0.5
0
0 100 200 300 400 500 600
1.5
y2
1
0.5
−0.5
0 100 200 300 400 500 600
comparison. The set-point responses are shown in Figure 7.6, exhibiting that
significant performance improvement has been achieved with the IMC design
and again its response is very close to that of the ideal IMC system. The
resulting ISE is calculated as 17.7, 51.5 and 17.0 for the IMC, BLT, and the
ideal IMC system, respectively. ♦
where
and gij0 (s) are strictly proper, stable scalar rational functions, and non-
negative Lij are the time delay associated with gij (s). Let the delay-free
part of the process be denoted by G0 = [gij0 ].
r + u y
G
-
+
Ĝ
-
+ +
Ĝ0
H(s) = G(s)G−1
0 (s)G0 (s)C(s)[I + G0 (s)C(s)]
−1
= G(s)G0 (s)−1 H0 (s).
The actual system performance could be quite poor due to the existence
of G(s)G−10 (s). For the special case where the delays of all the elements in
each row of the transfer matrix are identical, the finite poles and zeros in
G(s)G−1 −1
0 (s) will all be cancelled. In this case, G(s)G0 (s) = diag e
−Lii s
and the system output is the delayed output of H0 (s). However, in general
this desired property is not preserved.
In order to overcome this problem and improve the performance of the
multivariable Smith predictor control system, we here propose a decoupling
Smith predictor control scheme depicted in Figure 7.8, where D(s) is a de-
coupler for G, Q(s) the decoupled process G(s)D(s), Q0 (s) is the same as
Q(s) except that all the delays are removed. Suppose that G(s)D(s) is de-
coupled, it is obvious that the Q(s) and Q0 (s) will be diagonal matrices. The
multivariable smith predictor design is then simplified to multiple single-loop
smith predictor designs for which various methods can be applied. This de-
coupling Smith scheme involves three parts, decoupler D(s), the decoupled
process Q and the primary controller C(s). Their design will be discussed in
the next subsection.
7.2.2 Design
|G|
GD = diag{ dii , i = 1, 2, · · · , m}, (7.26)
Gii
Gij
dji = dii := ψji dii , ∀i, j ∈ m, j = i. (7.27)
Gii
where m = {1, 2, · · · , m}. We may adopt the procedure in Section 1 to design
D. But the present case has a separate C(s) which will take care of all control
issues except decoupling which is the only duty of D(s). Thus, we wish to
get a simplest decoupling D(s) with minimum calculations. Let θji as the
smallest non-negative number such that ψji e−θji s does not have prediction
for all j ∈ m. Let θi = max θji . By choosing dii = e−θi s in (7.27), it is
j∈m,j=i
obvious that dji will have no predictions.
258 7. Delay Systems
Each individual cii (s) is designed with respect to the delay free part qii0 of
qii such that closed-loop system formed by cii (s) and qii0 has the desired
performance. This is a number of SISO problems and deserves no further
discussions.
ũ = DCe, (7.31)
e = r − y + (Q − Q0 )Ce, (7.32)
7.2.4 Simulation
Several simulation examples are now given to show the effectiveness of the
decoupling Smith control.
7.2 The Smith Scheme 261
The diagonal elements of decoupler D(s) are chosen as d11 (s) = 1 and
d22 (s) = 1 since the ψ12 = g12 /g11 and ψ21 = g21 /g22 in (7.27) have no pre-
dictions. Model reduction yields the off-diagonal elements of the decoupler
as
(24.66s + 1.48)e−2s
d12 = ,
21s + 1
and
(4.90s + 0.34)e−4s
d21 = ,
10.9s + 1
and the decoupler is formed as
(24.66s+1.48)e−2s
1 21s+1
D(s) = .
(4.90s+0.34)e−4s
10.9s+1 1
The application of the model reduction to gii (s), the diagonal elements of
G(s)D(s), produces
6.37e−0.92s
g11 = ,
0.61s2 + 5.45s + 1
and
−9.65e−3.31s
g22 = ,
4.59s + 1
giving
6.37e−0.92s
0.61s2 +5.45s+1 0
Q(s) =
−9.65e−3.31s
0 4.59s+1 ,
and
6.37
0.61s2 +5.45s+1 0
Q0 (s) = −9.65
.
0 4.59s+1
0.066
0.23 + s − 0.09s 0
C(s) = 0.048
.
0 −0.14 − s + 0.097s
The set-point response is shown in Figure 7.11. For comparison, the Jerome’s
multivariable Smith predictor scheme is used with the controller settings
from Jerome and Ray (1986). It can be seen that the responses of the de-
coupling Smith control are much better and the decoupling is almost per-
fect. ♦
Example 7.2.2. Wardle and Wood (1969) give the following transfer function
matrix model for an industrial distillation column:
−6s −12s
0.126e −0.101e
60s+1 (45s+1)(48s+1)
G(s) = .
0.094e−8s −0.12e−8s
38s+1 35s+1
For this process, the diagonal elements of the decoupler D(s) are chosen as
d11 (s) = 1 and d22 (s) = 1 since the ψ12 = g12 /g11 and ψ21 = g21 /g22 have no
predictions. Model reduction yields
(48.09s+0.802)e−6s
1 2160s2 +93s+1
D(s) = .
47.72s2 +28.78s+0.783
66.14s2 +39.74s+1 1
7.2 The Smith Scheme 263
and Q0 (s) is readily obtained by taking the delay free part of Q(s). The
primary controller is then designed as
205.4 + 11.1
s − 47.75s 0
C(s) = .
0 −73.0 − 7.02
s − 17.2s
For comparison, the BLT tuning method (Luyben, 1986) is used here. The
set-point responses are shown in Figure 7.12, exhibiting that significant
performance improvement has been achieved with the decoupling Smith
design. ♦
Example 7.2.3. Consider the following system in Jerome and Ray (1986)
−2s −4s
(1−s)e 0.5(1−s)e
s2 +1.5s+1 (2s+1)(3s+1)
G(s) = .
0.33(1−s)e−6s (1−s)e−3s
(4s+1)(5s+1) 4s2 +6s+1
2
+0.75s+0.5)e−2s
1 − (0.5s 6s2 +5s+1
D(s) = 2
.
+1.98s+0.33)e−3s
− (1.32s 20s2 +9s+1 1
(16.9s2 +10.7+1)e−4.54s
17.4s2 +9.77s+1.2 0
Q(s) = ,
(5.18s+1)e−4.83s
0 26.79s2 +9.72s+1.2
and
0.37
−0.22 + s + 0.11s 0
C(s) = 0.16
.
0 1.79 + s − 0.90s
Fig. 7.14. Control System Robustness for Example 7.2.3 (— Proposed; - - - Jerome)
The first time delay compensation scheme was proposed by Smith (1957) for
SISO plants and is named after him. This scheme was extended to multivari-
able systems with single delay by Alevisakis and Seborg (1973),Alevisakis and
Seborg (1974) and with multi-delays by Ogunnaike and Ray (1979),Ogun-
naike et al. (1983). Due to multivariable interactions, the performance of such
MIMO Smith schemes might sometimes be poor ((Garcia and Morari, 1985).
Jerome and Ray (1986) proposed a different version of multivariable Smith
predictor control to improve overall performance. However, their design of
the primary controller is based on a transfer matrix with time delays and is
difficult to carry out. This also contradicts the Smith’s philosophy of delay
removal from closed-loop characteristic equation for ease of stable design. The
266 7. Delay Systems
work presented in Section 2 is based on Wang et al. (2000) and can achieve
better performance.
The IMC was introduced by Garcia and Morari (1982) and studied thor-
oughly in Morari and Zafiriou (1989) as a powerful control design strategy for
linear systems. In principle, if a multivariable delay model, Ĝ(s), is factor-
ized as Ĝ(s) = Ĝ+ (s)Ĝ− (s) such that Ĝ+ (s) is diagonal and contains all the
time delays and non-minimum phase zeros of Ĝ(s), then the IMC controller
can be designed, like the scalar case, as K(s) = {Ĝ− (s)}−1 with a possible
filter appended to it, and the performance improvement of the resultant IMC
controller over various versions of the multivariable Smith predictor schemes
can be expected (Garcia and Morari, 1985). For a scalar transfer function, it
is quite easy to obtain Ĝ+ (s). However, it becomes much more difficult for a
transfer matrix with multi-delays. The factorization is affected not only by
the time delays in individual elements but also by their distributions within
the transfer function matrix, and a non-minimum phase zero is not related to
that of elements of the transfer matrix at all (Holt and Morari, 1985b; Holt
and Morari, 1985a). The systematic method in Section 1 can effectively deal
with such a problem and it is from Zhang (1999).
8. Near-Decoupling
ẋ = A0 x + B0 u, y = C0 x,
where
Q.-G. Wang: Decoupling Control, LNCIS 285, pp. 267-292, 2003.
Springer-Verlag Berlin Heidelberg 2003
268 8. Near-Decoupling
−1 0 0 1 0
100
A0 = 0 −2 0 , B0 = 2 3 , C0 = .
111
0 0 −3 −3 −3
u = Kx + Gv,
where
" # " #
1 0 0 1 0
K= , G= .
− 53 − 43 −3 − 43 1
3
Now suppose that there are some parameter perturbations in matrix A0 , i.e.,
suppose
−1 + a1 0 0
A0 = 0 −2 + a2 0 ,
0 0 −3 + a3
Under the same decoupling law, the closed-loop transfer matrix of the per-
turbed system is as follows
" 1 #
s−a1 0
Hcl (s) = (a3 −2a2 +a1 )s+a1 a2 +a2 a3 −2a1 a3 −a1 +4a2 −3a3 a −a +1
.
2 3
(s−a1 )(s2 −(a2 +a3 )s+6a2 −6a3 +a2 a3 ) s2 −(a2 +a3 )s+6a2 −6a3 +a2 a3
a1 = a3 − 2,
(8.1)
a2 = a3 − 1.
In this section, we first define the concept of near-decoupling and then develop
a method to find solutions for near-decoupling controllers using state feedback
for systems with no perturbations, which will lay a basis for the developments
of the sections to be followed. Let us consider a linear time-invariant system
(A, B, C, D) described by
ẋ = Ax + Bu,
(8.2)
y = Cx + Du,
C1 D11 D12 · · · D1p
C2 D21 D22 · · · D2p
B = [B1 , B2 , · · · , Bp ], C = . , D= . .. .. . .
.. .. . . ..
Cp Dp1 Dp2 · · · Dpp
with
u = Kx + v (8.5)
such that the closed loop system (8.2)-(8.5) is near-decoupled for the new
input-output pairs (vi , yi ), i = 1, 2, . . . , p, in the sense of Definition 8.2.1.
Lemma 8.2.1. (Boyd et al., 1994) Denote by H(s) the transfer matrix of a
system (A, B, C, D). Then the condition
kHk∞ < γ
To deal with the requirement (8.3), let us first establish the following
result.
Lemma 8.2.2. Consider system (8.2) with transfer matrix H(s). Then the
inequality
holds.
Proof. The only if part. First notice, in terms of Parseval’s theorem, the fact
that inequality (8.7) is equivalent to the following inequality:
Z +∞ Z +∞
∗ 2
Y (jω)Y (jω)dω ≥ β U ∗ (jω)U (jω)dω ∀U ∈ H2 , (8.8)
−∞ −∞
where U (jω) and Y (jω) denote the Fourier’s transform of u(t) and y(t) re-
spectively. For a brief introduction on the spaces L2 (−∞, +∞) and H2 , read-
ers is referred to (Zhou et al., 1996). Inequality (8.6) means that
Thus we have
Z +∞ Z +∞
1
y T (t)y(t)dt = Y ∗ (jω)Y (jω)dω
−∞ 2π −∞
Z +∞
1
= U ∗ (jω)H ∗ (jω)H(jω)U (jω)dω
2π −∞
Z
β 2 +∞ ∗
≥ U (jω)U (jω)dω
2π −∞
Z +∞
= β2 uT (t)u(t)dt.
−∞
272 8. Near-Decoupling
where
σ1 (jω) 0 ··· 0
0 σ2 (jω) · · · 0
.. .
.. . .. .
..
Λ(jω) = .
0 0 · · · σm (jω)
0 0 ··· 0
Notice that the last row of zeros actually includes l − m rows of zeros. Now
we have
|σ1 (jω)|2 0 ··· 0
0 |σ (jω)|2
· ·· 0
2 ∗
H ∗ (jω)H(jω) = ER (jω) .. .. . .. ER (jω)
. . .. .
0 0 · · · |σm (jω)|2
Notice that there might be many such closed intervals and the length of
the such interval might be infinite. In this case, we just choose one of such
intervals and limit its length to be finite. Define em = [0 · · · 0 1]T ∈ Rm .
Now we choose the signal U (jω) as follows
ER (jω)em when ω ∈ I,
U (jω) =
0 otherwise.
and
8.2 Near-Decoupling for Unperturbed Systems: State Feedback 273
Z +∞
Y ∗ (jω)Y (jω)dω
−∞
Z
= ∗
eTm ER (jω)ER (jω)Λ∗ (jω)Λ(jω)ER
∗
(jω)ER (jω)em dω
I
Z
= |σm (jω)|2 dω
I
Lemma 8.2.3. Denote by H(s) the transfer matrix of system (8.2). If there
exists a matrix P > 0 such that
−AT P − P A + C T C −P B + C T D
≥ 0, (8.11)
−B T P + DT C −β 2 I + DT D
holds.
Proof. Let V (x) = −xT P x. Consider the motion of system (8.2) with zero
initial condition, i.e., x(0) = 0. If one can prove that
dV (x)
+ y T (t)y(t) − β 2 uT (t)u(t) ≥ 0, ∀u ∈ L2 (−∞, +∞), (8.13)
dt
then one has
Z +∞
[y T (t)y(t) − β 2 uT (t)u(t)]dt
0
Z +∞
dV (x)
≥− dt = xT (+∞)P x(+∞) − xT (0)P x(0) ≥ 0.
0 dt
Thus
Z +∞ Z +∞
T 2
y (t)y(t)dt ≥ β uT (t)u(t)dt, ∀u ∈ L2 (−∞, +∞).
0 0
From Lemma 8.2.2, we can therefore conclude that inequality (8.12) holds
true.
274 8. Near-Decoupling
Thus it is clear that if inequality (8.11) holds, inequalities (8.13) and then
(8.12) hold too.
Corollary 8.2.1. For system (8.2), if there exists a matrix P > 0 such that
−AT P − P A −P B + C T D
≥ 0,
−B T P + DT C −β 2 I + DT D
Remark 8.2.1. From Lemma 8.2.3 it can be seen that the parameter β is
upper bounded by
q
β ≤ λmin (DT D).
Based upon the above development, we can establish the following theo-
rem.
Theorem 8.2.1. For system (8.2), if there are matrices Q > 0 and F such
that the following LMIs
−QAT − AQ − F T B T − BF −Bi + (QCiT + F T DiT )Dii
> 0,
−BiT + DiiT
(Ci Q + Di F ) −βi2 I + Dii
T
Dii
∀i ∈ {1, 2, . . . , m}; (8.14)
QAT + AQ + F T B T + BF Bj QCiT + F T DiT
BjT −γI T
Dij < 0,
Ci Q + Di F Dij −γI
∀i, j ∈ {1, 2, . . . , m}, i 6= j. (8.15)
hold, then system (8.2) is nearly state-feedback decouplable. The state-feedback
law is given by
u = Kx + v, K = F Q−1 (8.16)
i.e., matrix inequalities (8.19) and (8.20) are equivalent to LMIs (8.14) and
(8.15), respectively. In regard to the stableness of the closed-loop system, we
notice that both (8.19) and (8.20) imply that (A+ BF )T P + P (A+ BK) < 0,
which further implies that the closed-loop system is stable. This completes
the proof.
Notice that the argument in the end of the above proof also applies to
Theorems 8.3.1, 8.4.1, and 8.5.1 given later on. Hence we will not repeat it
in the proof of these theorems.
In the decoupling controller design for practical plants, the magnitude of
parameters βi , i = 1, 2, . . . , m, can be readily estimated from Remark 8.2.1,
but it is hard to infer the magnitude of parameter γ. Therefore we formulate
the near-decoupling controller design (NDCD) in the case of state-feedback
(SF) as the following optimization problem.
276 8. Near-Decoupling
Problem NDCD-SF:
minimize γ
subject to LMIs Q > 0, (8.14), (8.15).
The above problem is a standard eigenvalue problem (EVP) in LMI lan-
guage (Boyd et al., 1994), hence its numerical solutions can be found through
LMI toolbox of Matlab very efficiently (Gahinet et al., 1995).
Now consider the case that there are some parameter perturbations in the
system matrix A. We assume that the perturbations are structured as follows
A = A0 + Al Ψ Ar , (8.21)
n1 ×n2
Ψ ∈ Ψ̆ := {Ψ ∈ R : kΨ k2 ≤ 1},
where A0 , a constant matrix, is the nominal parameter of A, Ψ denotes the
uncertainty, Al and Ar , constant matrices with appropriate dimensions, are
used to describe how the uncertainty Ψ enters the system matrix A. Notice
that any uncertainties with bounded norm can be described by the form
(8.21).
Theorem 8.3.1. Consider uncertain system (8.2) with matrix A described
by (8.21). If there are matrices Q > 0, F and numbers εij > 0 (i, j =
1, 2, . . . , p) such that the following LMIs
−QAT0 − A0 Q − F T B T − BF − εii Al ATl −Bi + (QCiT + F T DiT )Dii QATr
T T
−Bi + Dii (Ci Q + Di F ) 2 T
−βi I + Dii Dii 0 > 0,
Ar Q 0 εii I
∀i ∈ {1, 2, . . . , m}, (8.22)
QAT0 + A0 Q + F T B T + BF + εij Al ATl Bj QCiT + F T DiT QATr
T T
Bj −γI Dij 0 < 0,
Ci Q + Di F Dij −γI 0
Ar Q 0 0 −εij I
∀i, j ∈ {1, 2, . . . , m}, i 6= j, (8.23)
by applying Schur complements (Boyd et al., 1994). Then for any two real
numbers ε1 and ε2 we will have
I Υ ε1 ΦTl
ε1 Φl −ε2 ΦTr ≥ 0. (8.25)
ΥT I −ε2 Φr
are satisfied for some positive numbers εij (i, j = 1, 2, . . . , m). It follows from
Schur complements that inequalities (8.28) and (8.29) are equivalent to LMIs
(8.22) and (8.23), respectively. Thus the proof is completed.
Similarly, we can formulate the robust near-decoupling controller design
(RNDCD) in the case of state-feedback as the following optimization problem.
Problem RNDCD-SF:
minimize γ
subject to LMIs Q > 0, (8.22), (8.23).
In both problems NDCD-SF and RNDCD-SF, the minimization proce-
dure will give a very large Q, making Q−1 ill-conditioned. Therefore for these
two problems, we impose the following additional constraint
Q < µI
i.e.,
ż = AK z + BK Cx,
(8.31)
u = CK z + DK Cx + v.
We suppose that
|DK D + I| 6= 0.
Our basic idea is to use the results already obtained hereto to develop a
method for the design of controller (8.30). To proceed, we write the closed-
loop system consisting of (8.2) and (8.31) as follows
ẋ A + BDK C BCK x B
ż = BK C AK z
+
0
v,
(8.32)
x
y = C + DDK C DCK + Dv.
z
8.4 Near-Decoupling: Dynamic Output Feedback 279
Let
A + BDK C BCK B
Ā = , B̄ = , C̄ = C + DDK C DCK , D̄ = D.
BK C AK 0
Then it follows from Lemma 8.2.1 and Corollary 8.2.1 that system (8.32) is
near-decoupled with performance requirement (8.17)-(8.18) being satisfied if
there exists a matrix P̄ > 0 such that the following matrix inequalities
−ĀT P̄ − P̄ Ā −P̄ B̄i + C̄iT D̄ii
> 0, ∀i ∈ {1, 2, . . . , m}, (8.33)
T
−B̄iT P̄ + D̄ii T
C̄i −βi2 I + D̄ii D̄ii
ĀT P̄ + P̄ Ā P̄ B̄j C̄iT
B̄jT P̄ −γI D̄ij T
< 0, ∀i, j ∈ {1, 2, . . . , m}, i 6= j (8.34)
C̄i D̄ij −γI
hold.
Due to the fact that the matrix Ā also includes the design parameters
AK , BK , CK and DK , (8.33) and (8.34) are bilinear matrix inequalities in the
unknowns P̄ , AK , BK , CK and DK , which is very difficulty to solve. Thanks
to the pioneer work by Chilali, Gahinet and Scherer (Chilali and Gahinet,
1996; Scherer et al., 1997), both (8.33) and (8.34) can be transformed into
two LMIs by taking a change of variables. In the following, we will use the
method developed in (Chilali and Gahinet, 1996; Scherer et al., 1997) to solve
our problem.
Let us partition P̄ and P̄ −1 as
Y N −1 X M
P̄ = , P̄ = , (8.35)
NT ? MT ?
M N T = I − XY. (8.36)
It will be clear later that X and Y , together with other matrices, are design
variables which make the required performance indices be satisfied, while M
and N are free parameters, the unique requirement for which is to satisfy
equation (8.36). Define
X I I Y
Π1 = , Π2 = .
MT 0 0 NT
It is clear that
P̄ Π1 = Π2 .
280 8. Near-Decoupling
−BiT + Dii T
(Ci X + Di Ĉ) −BiT Y + DiiT
(Ci + Di D̂C) −βi2 I + Dii
T
Dii
T T
Al Ψ Ar X + X(Al Ψ Ar ) Al Ψ Ar + (Y Al Ψ Ar X) 0
− (Al Ψ Ar )T + Y Al Ψ Ar X Y Al Ψ Ar + (Al Ψ Ar )T Y 0
0 0 0
> 0, ∀i ∈ {1, 2, . . . , m}, ∀Ψ ∈ Ψ̆ , (8.48)
8.5 Robust Near-Decoupling: Dynamic Output Feedback 283
AX + XAT + B Ĉ + Ĉ T B T A + B D̂C + ÂT Bj (Ci X + Di Ĉ)T
(A + B D̂C)T + Â Y A + AT Y + B̂C + C T B̂ T T
Y Bj (Ci + Di D̂C)
BjT BjT Y −γI T
Dij
Ci X + Di Ĉ Ci + Di D̂C Dij −γI
A0 X + XAT0 + B Ĉ + Ĉ T B T A0 + B D̂C + ÂT0 Bj (Ci X + Di Ĉ)T
(A0 + B D̂C)T + Â0 Y A0 + AT0 Y + B̂C + C T B̂ T T
Y Bj (Ci + Di D̂C)
=
BjT BjT Y −γI T
Dij
Ci X + Di Ĉ Ci + Di D̂C Dij −γI
Al Ψ Ar X + X(Al Ψ Ar )T Al Ψ Ar + (Y Al Ψ Ar X)T 0 0
(Al Ψ Ar )T + Y Al Ψ Ar X Y Al Ψ Ar + (Al Ψ Ar )T Y 0 0
+
0 0 0 0
0 0 00
< 0, ∀i, j ∈ {1, 2, . . . , m}, i 6= j, ∀Ψ ∈ Ψ̆ (8.49)
together with (8.47) hold, system (8.2) is near-decoupled and the closed-loop
performance indices are given by (8.17)-(8.18). In view of Lemma 8.3.1, we
have
Al Ψ Ar X + X(Al Ψ Ar )T Al Ψ Ar + (Y Al Ψ Ar X)T 0
(Al Ψ Ar )T + Y Al Ψ Ar X Y Al Ψ Ar + (Al Ψ Ar )T Y 0
0 0 0
T
Al Ψ Ar X Al Ψ Ar 0 Al Ψ Ar X Al Ψ Ar 0
= Y Al Ψ Ar X Y Al Ψ Ar 0 + Y Al Ψ Ar X Y Al Ψ Ar 0
0 0 0 0 0 0
T
Al Al
= Y Al Ψ Ar X Ar 0 + Y Al Ψ Ar X Ar 0
0 0
Al XATr
−1
≤ ε3 Y Al ATl ATl Y 0 + ε3 ATr Ar X Ar 0 , ∀ε3 ∈ (0, +∞)
0 0
and
Al Ψ Ar X + X(Al Ψ Ar )T Al Ψ Ar + (Y Al Ψ Ar X)T 0 0
T T
(Al Ψ Ar ) + Y Al Ψ Ar X Y Al Ψ Ar + (Al Ψ Ar ) Y 0 0
0 0 00
0 0 00
Al XATr
Y A l
Ar
T
≤ ε−1 T T
4 0 Al Al Y 0 0 + ε4 Ar X Ar 0 0 , ∀ε4 ∈ (0, +∞).
0
0 0
Therefore, if the following inequalities
−A0 X − XAT0 − B Ĉ − Ĉ T B T −A0 − B D̂C − ÂT0 −Bi + (Ci X + Di Ĉ)T Dii
T
−(A0 + B D̂C) − Â0 −Y A0 − A0 Y − B̂C − C B̂ −Y Bi + (Ci + Di D̂C) Dii
T T T T
−BiT + DiiT
(Ci X + Di Ĉ) −BiT Y + Dii
T
(Ci + Di D̂C) −βi2 I + Dii
T
Dii
T
Al XA r
−ε−1
3
Y Al ATl ATl Y 0 − ε3 ATr Ar X Ar 0
0 0
> 0, ∀i ∈ {1, 2, . . . , m}, (8.50)
284 8. Near-Decoupling
A0 X + XAT0 + B Ĉ + Ĉ T B T A0 + B D̂C + ÂT0 Bj (Ci X + Di Ĉ)T
T
(A0 + B D̂C) + Â0 Y A0 + AT0 Y + B̂C + C T B̂ T Y Bj (Ci + Di D̂C)T
BjT BjT Y −γI T
Dij
Ci X + Di Ĉ Ci + Di D̂C Dij −γI
Al XATr
Y Al T T ATr
+ε−1
4 0 Al Al Y 0 0 + ε4 0 Ar X Ar 0 0
0 0
< 0, ∀i, j ∈ {1, 2, . . . , m}, i 6= j (8.51)
calculate ε∗3 and ε∗4 , which are then used as the initial values for ε3 and ε4
respectively in the sweeping procedure.
Based upon the above discussion, we can adopt the following procedure to
solve robust near-decoupling controller design (RNDCD) problem for output
feedback case.
ẋ = Ax + Bu, y = Cx + Du
with
−1 + ∆ 0 0 1 0
A= 0 −2 + 2∆ 0 , B = 2 3,
0 0 −3 + 3∆ −3 −3
100 10
C= , D= ,
111 01
286 8. Near-Decoupling
The Bode diagrams for this system (nominal part) are illustrated in Fig. 8.1.
Using the methods developed in the preceding sections, we can obtain the
near-decoupling controllers for SF and OF cases and achievable performance,
as summarized in Table 8.1.
β1 = β2 = 0.95,
−1.0009 0.0006 0.0009
RNDCD-SF γ = 4.43 × 10−4 K =
µ = 104 −0.9993 −1.0001 −0.9998
0.0103 6.6284 −6.6386
AK = 0.0159 6.6257 −6.6416 × 108
0.0159 6.6257 −6.6416
−1.0485 −2.1362
BK = −0.5438 −1.8575 × 108
NDCD-OF β1 = β2 = 0.95 γ = 6.7166 × 10−4 −0.5438 −0.5748
η = 104
0.0223 −0.0105 −0.0118
CK =
7.2e{−9} 2.0e{−5} −2.0e{−5}
DK = D
−2.2333 −1.2025 −0.5697
AK = −0.0930 −2.1850 −2.7781 × 106
−0.0930 −2.1850 −2.7781
−5.9078 −2.4216
β1 = β2 = 0.95 γ = 0.001 at BK = 1.7431 −4.1658 × 108
NDCD-ROF 1.7431 −4.1658
η = 10 ε3 = ε4 = 0.001
0.0063 −0.0002 −0.0030
CK =
0.0031 0.0114 0.0111
DK = D
The Bode diagrams for the closed-loop system (we choose ∆ = 0.1) are
illustrated in Figs. 8.2, 8.3, 8.4 and 8.5 for the cases NDCD-SF, RNDCD-SF,
NDCD-OF and RNDCD-OF respectively.
8.6 A Numerical Example 287
Comparing Table 8.1 and Figs. 8.2, 8.3, 8.4 and 8.5, we can see that the
performance indices guaranteed by our methods are somewhat conservative,
especially for the case where there are parameter perturbations. This is not
strange since the performance index γ shown in Table 8.1 is guaranteed for
all possible uncertain models.
2.2 1
0.8
2
0.6
1.8
0.4
0.2
1.6
|H11|
|H12|
1.4
−0.2
−0.4
1.2
−0.6
1
−0.8
0.8 −1
−2 −1 0 1 2 3 −2 −1 0 1 2 3
10 10 10 10 10 10 10 10 10 10 10 10
Frequency (rad/sec) Frequency (rad/sec)
1.2 1.6
1.5
1
1.4
0.8
1.3
|H21|
|H22|
0.6
1.2
0.4
1.1
0.2
1
0 0.9
−2 −1 0 1 2 3 −2 −1 0 1 2 3
10 10 10 10 10 10 10 10 10 10 10 10
Frequency (rad/sec) Frequency (rad/sec)
−4
1.0001 x 10
4
1 3.5
0.9999 3
2.5
0.9998
|
cl,11
|Hcl,12|
2
|H
0.9997
1.5
0.9996
1
0.9995
0.5
0.9994 0
−2 −1 0 1 2 3 −2 −1 0 1 2 3
10 10 10 10 10 10 10 10 10 10 10 10
Frequency (rad/sec) Frequency (rad/sec)
−4 −3
x 10 x 10
3.5 10.05
10
3
9.95
2.5
9.9
|Hcl,22|−0.99
2
|Hcl,21|
9.85
1.5
9.8
1
9.75
0.5
9.7
0 9.65
−2 −1 0 1 2 3 −2 −1 0 1 2 3
10 10 10 10 10 10 10 10 10 10 10 10
Frequency (rad/sec) Frequency (rad/sec)
−4
1.0001 x 10
2
1 1.8
1.6
0.9999
1.4
0.9998
1.2
|
cl,11
|Hcl,12|
0.9997
1
|H
0.9996 0.8
0.6
0.9995
0.4
0.9994
0.2
0.9993 0
−2 −1 0 1 2 3 −2 −1 0 1 2 3
10 10 10 10 10 10 10 10 10 10 10 10
Frequency (rad/sec) Frequency (rad/sec)
−4 −3
x 10 x 10
4 10.05
3.5 10
3 9.95
2.5 9.9
|Hcl,22|−0.99
|Hcl,21|
2 9.85
1.5 9.8
1 9.75
0.5 9.7
0 9.65
−2 −1 0 1 2 3 −2 −1 0 1 2 3
10 10 10 10 10 10 10 10 10 10 10 10
Frequency (rad/sec) Frequency (rad/sec)
−6 −7
x 10 x 10
3.9 1.4
3.85
1.2
3.8
1
3.75
0.8
|Hcl,11|−1.0
|Hcl,12|
3.7
0.6
3.65
0.4
3.6
0.2
3.55
3.5 0
−2 −1 0 1 2 3 −2 −1 0 1 2 3
10 10 10 10 10 10 10 10 10 10 10 10
Frequency (rad/sec) Frequency (rad/sec)
−7 −5
x 10 x 10
1.4 10.004
10.002
1.2
10
1
9.998
|−0.9999
0.8
|Hcl,21|
9.996
cl,22
0.6
|H
9.994
0.4
9.992
0.2
9.99
0 9.988
−2 −1 0 1 2 3 −2 −1 0 1 2 3
10 10 10 10 10 10 10 10 10 10 10 10
Frequency (rad/sec) Frequency (rad/sec)
u = Kx + Gv.
8.7 Concluding Remarks 291
−7 −8
x 10 x 10
10.4 1.4
1.2
10.2
1
10
0.8
|Hcl,11|−1.0
|Hcl,12|
9.8
0.6
9.6
0.4
9.4
0.2
9.2 0
−2 −1 0 1 2 3 −2 −1 0 1 2 3
10 10 10 10 10 10 10 10 10 10 10 10
Frequency (rad/sec) Frequency (rad/sec)
−7 −7
x 10 x 10
5.5 4
5 3.5
4.5 3
4 2.5
|Hcl,22|−1.0
|Hcl,21|
3.5 2
3 1.5
2.5 1
2 0.5
1.5 0
−2 −1 0 1 2 3 −2 −1 0 1 2 3
10 10 10 10 10 10 10 10 10 10 10 10
Frequency (rad/sec) Frequency (rad/sec)
A literature survey will reveal that nominal decoupling problem has now
been well studied. In sharp contrast to other control topics where robustness
has been extensively investigated, robust decoupling problem for uncertain
systems has rarely been reported in the literature. This phenomena is un-
satisfactory since parameter perturbations or deviations from their nominal
values are inevitable in real plants, especially in industrial process control.
Intuitively, exact decoupling is very difficult to achieve owing to the presence
of plant parameter perturbations. In practice, it is often the case that one first
designs a decoupling controller (decoupler) for the nominal model of the real
plant, put the decoupler as the inner loop controller and then on-line tune
the outer loop controller to make the overall system performance be roughly
satisfied. Reference (Wu et al., 1998) describes such a procedure. In robot
control systems, the calculated torque method combined with gain schedul-
ing (Ito and Shiraishi, 1997) also belongs to this approach. It is believed that
near or robust decoupling problem deserves more attention from researchers.
9. Dynamic Disturbance Decoupling
dimensions represent the plant model, sensor model and disturbance channel
model, respectively. Suppose that both ym and d are accessible. A general lin-
ear compensator which makes use of all the available information is described
by
u(s) = −Gf (s)d(s) − Gb (s)ym (s), (9.3)
where Gf and Gb are proper rational matrices to be designed for distur-
bance decoupling. This compensator can be thought as a combination of a
disturbance feedforward part and a output feedback part. The overall control
system is depicted in Figure 9.1. For the sake of simplicity the intermediate
variable s will be dropped as long as the omission causes no confusion.
It is well known that (9.1)∼(9.3) can equivalently be described by poly-
nomial matrix fraction representations:
Ay = B u + C d, (9.4)
Fξ= y, (9.5)
ym = E ξ, (9.6)
Pu= −Q d − R ym , (9.7)
where two pairs of polynomial matrices, A and [B C], and P and [−Q −R],
are both left coprime such that A−1 [B C] = [Gu Gd ] and P −1 [−Q −R] =
[−Gf − Gb ]; and F and E are right coprime such that EF −1 = Gm .
Combining (9.4)∼(9.7) yields
9.1 Measurable Disturbance: Feedforward Control 295
−B 0 A u C
0 F −I ξ = 0 d. (9.8)
P RE 0 y −Q
Let
−B 0 A
W = 0 F −I , (9.9)
P RE 0
then the transfer function matrix between y and d is
Gyd = [0 0 I]W −1 [C T 0 − QT ]T . (9.10)
We can now formulate the disturbance decoupling problem with stability
as follows. (DDP): given a controlled system described by (9.1) and (9.2)
with Gu , Gd and Gm being proper rational matrices, find the conditions
under which a compensator in the form of (9.3) with Gf and Gb being proper
rational matrices will make (i) the resulting system stable, i.e., all the roots
of the determinant det(W ) lie in C− , the open left half of complex plane; and
(ii) the transfer matrix between y and d equals zero, i.e., Gyd = 0.
9.1.1 Solvability
Our strategy to resolve (DDP) is first to determine the set of proper compen-
sators which stabilize a controlled system. Within these stabilizable systems
and their compensators, the conditions for zeroing disturbance are then de-
rived.
Let G1 and G2 be rational matrices with the same number of rows, we
can factorize them as
G1 = D1−1 N1 , G2 = D2−1 N2 , (9.11)
Lemma 9.1.1. With the above notations, D̃22 is a right divisor of any great-
−1
est common left divisor of N31 and D3 , where D̃22 D̃1 is a left coprime frac-
−1
tion of D1 D22 .
296 9. Dynamic Disturbance Decoupling
d ~
Gd
~
Gf
- + y
~ u ~
Gb ( D +f ) −1 Gu ( Dd+ ) −1
- +
ym
Gm
Remark 9.1.2. To know if Gu has the unstable model of Gd , one only needs
to check if Du Dd−1 is stable, which is much easier to perform without a need
to factorize Dd into Dd− Dd+ . It is also sufficient if Dd Gu is stable.
Consider now (DDP) under the stability conditions given in Theorem
9.1.1. From Figure 9.1, we have
Gyd = [I + Gu Gb Gm ]−1 [Gd − Gu Gf ].
Gyd = 0 is equivalent to
Gu Gf = Gd . (9.25)
Lemma 9.1.2. If the conditions of Theorem 9.1.1 are satisfied and Gf meets
(9.25), then Gf is stable.
Proof. Assume that Gf satisfies (9.25) but is not stable. With the notations
as in Theorem 9.1.1, premultiplying (9.25) by Dd+ gives
(D̄u )−1 Nu Gf = G̃d ,
where G̃d = (Dd− )−1 Nd and D̄u = Du (Dd+ )−1 with D̄u being a polynomial
matrix due to Gu having the unstable model of Gd . It follows from the sta-
bility of G̃d and coprimeness of Nu and D̄u that Nu · Gf must has RHP
pole-zero cancellations. Denote by (Df+ )−1 such a cancelled unstable model
of Gf . Then Gf = (Df+ )−1 G̃f , σ(Df+ ) ⊂ C+ , and Nu (Df+ )−1 is a poly-
nomial matrix. Since Gb has the unstable model of Gf , it must have the
form Gb = (Df+ )−1 G̃b . This results in the RHP pole-zero cancellations in
Gu · Gb = Du−1 Nu (Df+ )−1 G̃b , which violates the condition (ii) of Theorem
9.1.1. Proof is therefore completed.
Theorem 9.1.2. (DDP) is solvable if and only if the following hold
(i) Gu has the unstable model of Gd ;
(ii) Gm · Gu has no RHP pole-zero cancellations; and
(iii) (9.25) has a proper stable solution Gf .
300 9. Dynamic Disturbance Decoupling
Proof. Assume that (DDP) is solvable, i.e., there are proper Gb and Gf such
that the system is stable and Gyd = 0. In view of Theorem 9.1.1, conditions
(i) and (ii) above must be satisfied in order to assure stability. In addition,
Gyd = 0 implies that Gf satisfies (9.25). It follows from Lemma 9.2 that Gf
must also be stable.
Conversely assume that the conditions (i) through (iii) are satisfied. Then
Gyd = 0 for the solution Gf . And there is a proper Gb which stabilizes Gm Gu .
Furthermore, such a Gb must have the unstable model of Gf due to Gf being
itself stable. Hence, by Theorem 9.1, the system is stable. And the above Gf
and Gb forms a solution of (DDP).
9.1.2 Synthesis
Theorem 9.1.3. Given Ḡ1 and Ḡ2 as above, (9.26) has a solution for G if
and only if (9.27) holds. If it is the case, all the solutions are given by the
following polynomial matrix fraction:
G = N D−1 , (9.30)
N = BW + N0 V, (9.31)
D = RV, (9.32)
where W and V are arbitrary polynomial matrices, and N0 satisfies
P̃1 N0 = P̃2 R. (9.33)
Furthermore,
(i) a solution is of minimal order if and only if V is unimodular. Without
loss of generality all the minimal solutions are given by (9.30) and (9.33)
with V = I;
302 9. Dynamic Disturbance Decoupling
Proof. It is clear that (9.27) is necessary for the existence of solutions. Con-
versely assume that it holds. Then (9.26) will have a solution if (9.28) has a
solution, which is guaranteed due to G1 having full rank. Indeed, replace G
in (9.28) by a right coprime polynomial matrix fraction,
G = N D−1 . (9.35)
Equation (9.28) can be rewritten as
P̃1 N = P̃2 D. (9.36)
−1
Premultiplying (9.36) by L yields
P̄1 N = P̄2 R−1 D. (9.37)
Because the columns of P̄1 are coprime by the definition of L, then there is
a polynomial matrix M such that P̄1 M = I. Setting N = M P̄2 and D = R,
then such N and D satisfy (9.37) and thus N D−1 is a solution of (9.28).
In order to search for a general solution of (9.36), it will be first shown
that if D is any solution of (9.36), it must have the form
D = RV, (9.38)
for some polynomial matrix V . Indeed, assume that D is a solution of (9.36),
it then satisfies (9.37). This implies that R−1 D must be a polynomial matrix
because P̄2 and R are coprime. In other words, D = RV for some polynomial
matrix V .
A general solution of (9.36) can be expressed as the sum of the general
solution of the associated homogeneous equation,
P̃1 N1 = 0, (9.39)
and a particular solution N2 of the inhomogeneous equation (9.36). If B is a
minimal base of the right null space of P̃1 , then it is well known (Forney 1975)
that the general solution N1 of (9.39) is N1 = BW , where W is a arbitrary
polynomial matrix. For the particular solution N2 of (9.36), set N2 = N0 V ,
where N0 satisfies
P̃1 N0 = P̃2 R, (9.40)
9.1 Measurable Disturbance: Feedforward Control 303
(s − 1)(s + 2) 0 (s + 2) 1
Du = , Nu = ,
0 (s − 1)(s + 1) (s + 1) 1
(s − 1)(s + 2) 0 1 0
Dd = , Nd = .
0 (s − 1)(s + 1)(s + 2) (s + 2) 1
Because
" #
1 0
Du Dd−1 = 1
0 s+2
Rank(P̃h1 ) = 2.
Thus a proper solution exists. Because P̃1 = P̃1 · I, the greatest common left
divisor of P̃1 can be taken as itself, i.e., L = P̃1 . And the minimal base B of
L is clearly zero. Hence
−1
−1 (s + 2) 1 10
L P̃2 =
0 1 11
" 1
# −1
0 − s+2 0 −1 1 0
= = := P̄2 R−1 . (9.42)
1 1 1 s + 2 0 s + 2
is solvable if and only if this Gf is proper and stable. The computations in the
example is thus not necessary, but is carried out to illustrate the procedure
for a general case.
For the feedback compensator Gb which stabilizes Gu there are many
well-known algorithms for its construction (Chapter 3). Therefore, specific
computations are omitted here.
In this section the disturbance decoupling problem with stability by dy-
namic measurement feedback and disturbance feedforward has been consid-
ered. The problem is solved in a polynomial matrix setting. The stability
conditions are established. The main observation is that for ensuring the sta-
bility, the transfer matrix of control channel and the feedback compensator
must have the unstable models of disturbance channel and of the feed-forward
compensator, respectively, and there should be no RHP pole-zero cancella-
tions between the plant and sensor. They are intuitively appealing and have
an explicit dynamic interpretation. Under such stability conditions it is then
shown that solvability of disturbance decoupling is equivalent to the exis-
tence of a proper stable solution for some rational matrix equation. It is also
shown that the construction of a compensator for (DDP) can be made into
two independent parts. One is the feedback compensator synthesis, which is
a well-known problem. Another is the feed-forward compensator synthesis,
which can be incorporated into a more general problem, namely the model
matching. A unified approach, which can simultaneously treat properness,
stability and minimality, is presented. This approach also gives an explicit
parametric form of all the solutions.
PC G
Yo = R+ D,
1 + PC 1 + PC
where the superscription o indicates the output response in the conventional
feedback system.
and
H(s) = P −1 (s). (9.46)
Let P (s) = P0 (s)e−Ls and its model P̂ (s) = P̂0 (s)e−L̂s , where P0 and P̂0
are delay-free rational transfer functions. Then, using the model, (9.45) and
(9.46) are replaced by
respectively. For a strictly proper P̂0 (s), its inverse is not physically realiz-
able, and nor is the pure predictor eL̂s . For the term eL̂s , the approximation
presented in Huang et al. (1990) as shown in Figure 9.5 is adopted here,
where the overall transfer function is
1 + V (s)
Gv (s) = .
1 + V (s)e−L̂s
If V (s) is of low-pass with high gain, then
V (s)
Gv (s) ≈ = eL̂s
V (s)e−L̂s
308 9. Dynamic Disturbance Decoupling
According to our study, the value of τv is suggested to be set as 0.1L̂ ∼ L̂. The
remaining parameter Kv is chosen as large as possible for high low-frequency
gain of Gv while preserving stability of Gv . The critical value Kvmax of Kv
at which the Nyquist curve of τvKs+1
v
e−L̂s touches −1 + j0 can be found from
q
Kvmax = 1 + ω ∗2 τv2 ,
tan−1 (ω ∗ τv ) + ω ∗ L̂ = π. (9.50)
For sufficient stability margin, Kv should be much smaller than Kvmax and
its default value is suggested to be Kvmax /3.
P̂0 (s) is usually strictly proper and P̂ 1(s) is then improper. A filter Q(s)
0
Q(s)
is introduced to make P̂0 (s)
proper and thus realizable. Q(s) should have at
least the same relative degree as that of P̂0 (s) to make Q(s)/P̂0 (s) proper,
and should be approximately 1 in the disturbance frequency range for a sat-
isfactory disturbance rejection performance. The following form is a possible
choice:
1
Q(s) = , τq > 0, (9.51)
(τq s + 1)n
initial value is set as 0.01Tp ∼ 0.1Tp and its tuning will be discussed below to
make the system stable. Other forms of Q(s) have been discussed by Umeno
and Hori (1991).
In view of the above development, H(s) is approximated by
The actual system to realize the proposed control scheme with such an ap-
proximation is depicted in Figure 9.6.
G (s )
r + + u + + y
C (s ) P0 ( s ) e − Ls
_ _
+
Pˆ0 ( s ) ˆ
e − Ls _
+
Q( s )
V (s )
+ _ + Pˆ ( s )
0
ˆ
e − Ls
Gv (s )
Q(s)
H(s) = ,
P̂ (s)
9.2.2 Analysis
Consider stability and performance of the control scheme in Figure 9.6. Let
a system consist of n subsystems with scalar transfer functions gi (s), i =
1, 2, · · · , n, and pi (s) be the characteristic polynomial of gi (s). Define
pc (s) = ∆(s)p0 (s), (9.53)
where ∆(s) is the system Q determinant as given by the Mason’s formula
n
(Mason, 1956) and p0 (s) = i=1 pi (s). It is shown (Chapter 3) that the sys-
tem is internally stable if pc (s) in (9.53) has all its roots in the open left-half
of the complex plane.
For the proposed scheme, write the transfer function of each block into
a coprime polynomial fraction with possible time delay, i.e., C(s) = ndcc(s) (s)
,
np (s) −Ls np̂ (s) −L̂s nv (s) nq (s)
P (s) = dp (s) e , P̂ (s) = dp̂ (s) e , V (s) = dv (s) , Q(s) = dq (s) and G(s) =
ng (s)
dg (s) ,
where nc (s), dc (s), np (s), dp (s), np̂ (s), dp̂ (s), nv (s), dv (s), nq (s) dq (s),
ng (s) and dg (s) are all polynomials. It follows from Figure 9.6 that
P0 P0
∆(s) = 1 + P C + QV e−Ls + Qe−Ls
P̂0 P̂0
− QV e−L̂s − Qe−L̂s + V e−L̂s + P CV e−L̂s ,
P0
= (1 + P C)(1 + V e−L̂s ) + (1 + V )Q( e−Ls − e−L̂s ), (9.54)
P̂0
p0 (s) = np̂ (s)dp (s)dp̂ (s)dv (s)dc (s)dq (s)dg (s). (9.55)
∆ˆ = (1 + P C)(1 + V e−Ls ),
so that
þ
ˆ 0 = {dp dc (1 + P C)} dv (1 + V e−Ls ) {np dp dq dg } .
p̂c = ∆p
The assumed stability of the conventional feedback system in Figure 9.3
means that dp dc (1 + P C) only has LHP roots. Under the assumptions that
the process is stable and without RHP zero, and G is stable, dp , dg and np
are then all stable. Therefore, it is concluded that the proposed scheme is
nominally stable, provided that stable Gv (equivalently dv (1 + V e−Ls )) and
Q (or dq ) are used.
Turn now to the case where there is a process/model mismatch:
P̂ (s) 6= P (s),
and the modelling error is described by
ü ü
ü ü
üP (jω) − P̂ (jω)ü < γ(jω).
9.2 Unmeasurable Disturbances: Disturbance Observers 311
where the perturbed term from the nominal case is δ(s) = (1+V )Q( P
P̂
0 −Ls
e −
0
e−L̂s ) and
ü ü
ü 1 + V (jω) ü
ü ü
|δ(jω)| = ü Q(jω)(P (jω) − P̂ (jω))ü
ü P̂0 (jω) ü
ü ü
ü 1 + V (jω) ü
ü ü
<ü Q(jω)ü γ(jω) := δ̄(jω).
ü P̂0 (jω) ü
Since the nominal stability has been proven, the Nyquist curve of ∆(s) ˆ
should have a correct number of encirclements of the origin, as required by
ˆ
the Nyquist stability criterion. One can superimpose the δ̄(jω) onto ∆(s)’s
Nyquist curve to form a Nyquist curve band, and check its stability.
Consider now the performance of the proposed method and its robustness.
With a stabilizing and tracking controller C, the process output response
yd (t) to a disturbance will settle down, i.e., yd (t) → 0 as t → ∞. Thus, the
widely used index in the time domain:
sZ
∞
kyd (t)k2 = yd2 (t)dt,
0
makes sense and can be used for performance evaluation. To have a better
disturbance rejection than the conventional feedback system requires
kyd (t)k2 6 kydo (t)k2 , (9.56)
where ydo (t) is the output response to the same disturbance if the conventional
feedback system is used. Applying the Parseval’s Identity (Spiegel, 1965):
Z ∞ Z ∞
1
yd2 (t)dt = |Yd (jω)|2 dω,
0 2π 0
to (9.56) gives
kYd k2 6 kYdo k2 , (9.57)
where
sZ
∞
kYd k2 , |Yd (jω)|2 dω,
0
sZ
∞
kYdo k2 , |Ydo (jω)|2 dω.
0
312 9. Dynamic Disturbance Decoupling
e−Ls
For illustration, let us consider the simple case with Gv = eLs , P = Tp s+1
and C(s) = kc + ksi . The disturbance is supposed to be D(s) = 1
s through a
1 1
dynamics G(s) = Td s+1 , and the filter is supposed to be Q(s) = τq s+1 . With
simple algebra, we evaluate
û û Z ∞
û GD û2
û û = A(ω)dω, (9.60)
û1 + PC û
2 0
where
A(ω)
Tp2 ω 2 + 1
=
(Td2 ω 2 + 1)[(−Tp ω 2 + ki cos ωL + kc ω sin ωL)2 + (ω + kc ω cos ωL − ki sin ωL)2 ]
> 0.
where ∗ is the standard convolution, and g1 and g2 are the impulse responses
of G1 (s) and G2 (s) respectively.
Proof. Suppose that x(t) has the period T and is expressed by its Fourier
series (Cartwright, 1990):
∞
X
x(t) = x0 + Ci cos(2πit/T + φi )
i=1
X∞
= x0 + Ci cos(ωi t + φi )
i=1
RT p
with x0 = T1 0 xdt, Ci = A2i + Bi2 , ωi = 2πi/T , and φi = − tan−1 (Bi /Ai ),
RT RT
where Ai = T2 0 cos(2πit/T )xdt and Bi = T2 0 sin(2πit/T )xdt. Having
such a signal transmitted through a stable G1 , the output is given by g1 (t) ∗
x(t). The steady state of the output is a periodic signal with the same period
T (Cartwright, 1990):
∞
X
(g1 ∗ x)s (t) = G1 (0)x0 + Ci |G1 (jωi )| cos(ωi t + φi + ∠G1 (jωi )).
i=1
Noting that
and
we then have
1 − e−k0 T s
Yd (s) = GD. (9.64)
1 − HF + P C + P H
Let η(t) the impulse response of G/(1 − HF + P C + P H). One obtains from
(9.64) that
G (s )
r + + u + + y
C (s ) P0 ( s ) e − Ls
_ _
+
Pˆ0 ( s ) ˆ
e − Ls _
Q( s )
e − Lh s
Pˆ0 ( s )
9.2.4 Simulation
e−5s
P (s) = .
10s + 1
The PI controller design method proposed in Zhuang and Atherton (1993) is
applied to give
1.01
C(s) = 1.01 + .
s
For the scheme in Figure 9.6 with perfect model assumption P̂ (s) = P (s),
take τv = 0.4L = 2. We obtain Kvmax = 1.38 from (9.50). Setting Kv =
0.46
Kvmax /3 = 0.46 yields V (s) = 2s+1 . The filter Q(s) is in the form of Q(s) =
1
τq s+1 , and τq is chosen as 0.1Tp = 1. A unit step change in the set-point is
made at t = 0, and a step disturbance of the size 1 is introduced through
1
G(s) = 20s+1 to the system at t = 50, the resulting responses are displayed in
Figure 9.8. Clearly, the disturbance observer control exhibits better capability
to reject load disturbance compared with the conventional controller.
To evaluate the scheme of Figure 9.7 for periodic disturbances, the fol-
lowing periodic disturbance is introduced at t = 50 as a separate test:
316 9. Dynamic Disturbance Decoupling
1
d
0.5
0
0 10 20 30 40 50 60 70 80 90 100
1.5
1
y
0.5
0
0 10 20 30 40 50 60 70 80 90 100
1
u
−1
0 10 20 30 40 50 60 70 80 90 100
1
0.5
0
d
−0.5
−1
0 10 20 30 40 50 60 70 80 90 100
1
y
−1
0 10 20 30 40 50 60 70 80 90 100
10
0
u
−10
−20
0 10 20 30 40 50 60 70 80 90 100
Fig. 9.9. Time Responses for Example 9.2.1 under Periodic Disturbance
(——Disturbance Observer; − − −Conventional Control)
In the development made so far, the process P (s) has been assumed to
be stable without any zeros in the RHP. However, the control scheme in
Figure 9.6 can be modified to Figure 9.12 so that it is applicable to unstable
processes and/or those with RHP zeros. In this case, the process P (s) is
factorized as
where P0− (s) represents the stable non-minimum-phase part, and P0+ (s) in-
cludes all RHP zeros and poles. Subsequently, P0− (s) is used in (9.47) and
(9.52) to design F (s) and H(s), respectively.
318 9. Dynamic Disturbance Decoupling
1
d
0.5
0
0 50 100 150
1.5
1
y
0.5
0
0 50 100 150
1.5
1
u
0.5
0
0 50 100 150
e−0.4s
P (s) = .
(s + 1)(s − 1)
We have so far considered the SISO case. The extension to the MIMO case is
straightforward. Suppose that the process P (s) is square, nonsingular, stable
and of minimum phase. The same scheme as in Figure 9.4 is exploited. It
then follows that
9.2 Unmeasurable Disturbances: Disturbance Observers 319
Im
1
−1
−2
−3
−4
−5
−6
−7
Re
−0.2 0 0.2 0.4 0.6 0.8 1 1.2
ˆ − − − δ)
Fig. 9.11. Nyquist Curve Banks For Example 9.2.2 (—— ∆;
Y = GD + P U, (9.66)
and
U = C(R − Y ) − H(Y − F U )
= CR + HF U − (C + H)(GD + P U ),
where p̃ij0 (s) are proper rational functions. But terms e−Lij s often happen to
be a pure predictor which is not physically realizable, that is, some Lij may be
negative. They will require the approximation shown in Figure 9.5, each for
one negative Lij . It is costly. To reduce the number of such approximations,
extract the largest possible pure prediction horizon L by
0, if all − Lij ≤ 0;
L= (9.70)
max{−L | − L > 0}, otherwise.
ij ij
1.05
0.95
0.9
y
0.85
0.8
0.75
0.7
0 50 100 150
0.5
−0.5
u
−1
−1.5
0 50 100 150
−0.4s
e
Fig. 9.13. Time responses for P (s) = (s+1)(s−1)
(—— Disturbance Observer; − − − Conventional Control)
∗ ∗ ∗
p̃110 (s)e−L11 s p̃120 (s)e−L12 s · · · p̃1m0 (s)e−L1m s
P̃ = diag {L, L, · · · , L} ··· ··· ··· ···
∗ ∗ ∗
p̃m10 (s)e−Lm1 s p̃m20 (s)e−Lm2 s · · · p̃mm0 (s)e−Lmm s
(9.71)
H(s) = P̃ (s)e−τ̃h s ,
where there are three pure predictors e2s , e2s and e4s in P̃ (s). According to
(9.70), we have
L = 4.
The Nyquist arrays of P −1 (s) and P̃ (s) are depicted in Figure 9.14, exhibiting
good matching of their frequency responses.
To approximate e4s , we set τv = L/2 = 2 and get Kvmax = 1.52. By setting
0.38
Kv = Kvmax /4 = 0.38, V (s) = 2s+1 is obtained. A unit step change is made
at t = 0 for the 1st set-point and at t = 200 for 2nd set-point, respectively, and
1
a step disturbance of size 1 is further introduced through Gd = 20s+1 to the
system at t = 400. The resulting output responses are shown in Figure 9.15.
It can be seen that with the add-on control to the conventional feedback
control, the system has achieved a better load disturbance rejection.
To evaluate the control scheme of Figure 9.7 for periodic disturbances,
the following periodic disturbance instead of a step disturbance is introduced
at t=400:
d(t) = sin(0.1t) + 0.3 sin(0.2t) + 0.1 sin(0.3t) + 0.1 sin(0.4t)
2π
with the period of T = 0.1 = 62.8 and G(s) = 1. τ̃h = k0 T = 62.8 is adopted
with k0 = 1. The system responses are shown in Figure 9.16. It is observed
9.3 Notes and References 323
0.5 0
−0.05
0.4
−0.1
0.3 −0.15
0.2 −0.2
−0.25
0.1
−0.3
0 −0.35
−1 −0.5 0 0.5 −0.2 0 0.2 0.4 0.6
0.05 0.6
0
0.4
−0.05
0.2
−0.1
0
−0.15
−0.2 −0.2
0 0.1 0.2 0.3 0.4 −0.2 −0.1 0 0.1 0.2 0.3
−1
Fig. 9.14. Nyquist Array of P (s) and P̃ (s) for Example 9.2.4
(—— P −1 (s), · · · P̃ (s))
1.5
1
d
0.5
0
0 100 200 300 400 500 600
1.5
1
y1
0.5
0
0 100 200 300 400 500 600
1.5
1
y2
0.5
0
0 100 200 300 400 500 600
d 0
−1
−2
0 100 200 300 400 500 600 700 800
1.5
y1 1
0.5
0
0 100 200 300 400 500 600 700 800
2
y2
0
0 100 200 300 400 500 600 700 800
Fig. 9.16. Control performance for Example 9.2.4 under Periodic Disturbance
(—— with Disturbance Observer, · · · without Disturbance Observer)
10. Asymptotic Disturbance Decoupling
~ d (s )
d Gd (s ) Gd (s )
~ +
r ~ r (s ) e(s ) u(s ) y (s )
Gr ( s ) Gc (s ) Gu (s )
+ +
-
y m (s )
Gm (s )
By the final value theorem of the Laplace transform, one can conclude that for
a stable system, the steady-state behavior of y(t) and e(t) is independent of
the stable parts, G̃− −
d and G̃r , of G̃d and G̃r . We can therefore assume without
loss of generality that any stable modes, which may be initially associated
with G̃d and G̃r , have been removed when defining G̃d and G̃r , in other
words, all the poles of both G̃d and G̃r can be assumed to lie in C+ .
We now formulate a general servocompensator problem with internal sta-
bility as follows. (SP): given a system in (10.1)-(10.3) and the disturbance
and reference in (10.4)-(10.5), as depicted in Figure 10.1 with Gu , Gm , G̃d ,
and G̃r all known and proper rational, Gu and Gm having no RHP pole-zero
cancellations and G̃d and G̃r being anti-stable (i.e., having the RHP poles
only), determine a proper compensator Gc such that the system of (10.1)-
(10.3) is internally stable and e(t) asymptotically tends to zero for any real
vectors d˜ and r̃ . In particular, the case of d˜ = 0 is called tracking problem
with internal stability (TP), and the case of r̃ = 0 and e replaced by y is
called regulation problem with internal stability (RP).
10.1.1 Preliminaries
Recall from Chapter 2 that two polynomial matrices, D ∈ Rm×m [s] and
N ∈ Rm×l [s], are left coprime (implicitly in C) if Rank[D(s) N (s)] =
m, f or all s ∈ C. This can be generalized as follows.
Definition 10.1.1. Two polynomial matrices, D ∈ Rm×m [s] and N ∈
Rm×l [s], are called left coprime in C+ , if
DN = N̄ D̄, (10.8)
10.1.2 Solvability
We will solve (RP) first, and subsequently apply its solution to (TP) and (SP).
Given Gu , Gm and G̃d , we can obtain polynomial fraction representations as
follows:
Gu =Du−1 Nu , (10.9)
−1
Gm =Dm Nm , (10.10)
−1
Du G̃d =Dd Nd , (10.11)
Nm Du−1 =D1−1 N1 , (10.12)
where all fractions on the right-hand sides are coprime and Dd−1 is anti-stable.
Theorem 10.1.1. (RP) is solvable if and only if
(i) Gu has the unstable model of Gd ;
(ii) N1 and Dd are right coprime in C+ ; and
(iii) Nu and Dd are externally skew-prime in C+ .
Proof. Sufficiency: The external skew-primeness of Nu and Dd in C+ means
that there are D̄d and N̄u such that Dd−1 N̄u = Nu D̄d−1 and both fractions
are coprime in C+ . Take
Gc = D̄d−1 Ḡc .
Write
Gm Gu D̄d−1 = [(D1 Dm )−1 N1 Nu ][D̄d−1 ].
Our problem formulation assumes that Gm · Gu has no RHP pole-zero can-
cellations. Internal stability will follow from Lemma 10.1.3 if Gm Gu D̄d−1 is
internally stabilizable, or equivalently, by Lemma 10.1.1, if (N1 Nu )D̄d−1 is
right coprime in C+ . Since both (N1 , Dd ) and (Nu , D̄d ) are right coprime in
C+ , there exist stable rational matrices X1 , Y1 , X2 , Y2 such that
X1 N1 + Y1 Dd = I,
X2 Nu + Y2 D̄d = I.
332 10. Asymptotic Disturbance Decoupling
Dd Nu = N̄u D̄d ,
where
−1
D̄d := D̄c1 Dc (10.17)
This implies that Nu and D̄d are right coprime in C+ , because otherwise
Dd−1 N̄u being equal to Nu D̄d−1 could not be coprime.
One notices
−1
Gm Gu Gc = D m (Nm Du−1 )(Nu Dc−1 )Nc
−1 −1
= Dm D1 N1 D̃c−1 Ñu Nc
= Dm D1 (N1 Dd−1 )D̃c1
−1 −1 −1
Ñu Nc .
Du−1 Dd−1 Nd = (Dd Du )−1 Nd , and the last representation is also a fraction
of G̃d . So there is a polynomial matrix L such that Dd Du = LD̃d . In the
light of Theorem 10.1, in order to solve (RP) we must set Gc = D̄d−1 Ḡc . One
sees
Gu Gc = Du−1 Nu D̄d−1 Ḡc
= Du−1 Dd−1 N̄u Ḡc
= (Dd Du )−1 N̄u Ḡc
= (D̃d−1 )(L−1 N̄u Ḡc ),
as shown in Figure 10.2. It is clear that when (RP) is solved, an exact du-
plication of the inverse of the denominator matrix of the exogenous system
must be present in the internal loop at the same summation junction as the
exogenous signals enter the loop. This is known as ”internal model principle”.
~ ~ ~
d N d ( s) Dd−1 ( s )
+
~ y (s )
−1
L N u Gc ( s ) Dd−1 ( s )
+
Gm (s )
Fig. 10.2. Internal Model Principle
To apply Theorem 10.1.1 to (TP). Figure 10.1 is redrawn into Figure 10.3.
Then, (TP) is obviously equivalent to (RP) with Gu , Gm and Gd replaced
by Gm Gu , I and I respectively. Let
−1
Gm Gu = Dmu Nmu ,
Dmu G̃r = Dr−1 Nr ,
where all fractions on the right hand sides are coprime. Noting that Gm Gu
and I cannot have any pole-zero cancellations and I has no unstable modes at
all so that they trivially meet condition (i) of Theorem 10.1, we obtain from
Theorem 10.1.1 the following corollary for solvability of asymptotic tracking
problem.
Corollary 10.1.1. (TP) is solvable if and only if Nmu and Dr are exter-
nally skew prime in C+ .
10.1 General Disturbances: Internal Model Principle 335
~
r ~ r
Gr ( s )
+
e(s )
Gc (s ) Gu (s ) Gm (s )
-
For the general (SP), we combine Theorem 10.1 and Corollary 10.1 as
follows.
Corollary 10.1.2. (SP) is solvable if and only if
(i) Gu has the unstable model of Gd ;
(ii) N1 and Dd are right coprime in C+ ;
(iii) Nu and Dd are externally skew-prime in C+ ; And
(iv) Nmu and Dr are externally skew-prime in C+ .
10.1.3 Algorithms
Based on Theorem 10.1.1, a compensator which solves (RP) can be con-
structed as follows. Given transfer matrices Gu , Gm and G̃d ,
(i) determine left coprime fractions in (10.9)–(10.12);
(ii) check the solvability conditions of Theorem 10.1. If they are met, con-
struct D̄d ; and
(iii) design a compensator Ḡc that stabilizes Gm Gu D̄d−1 and meets other
control requirements if any.
Lemma 10.1.5. For nonsingular D and full row rank N with σ(D) ∩
σ(N ) = ∅, the Smith form of DN is the product of those of D and N , that
is, S3 = S1 S2 .
Proof. From (10.19) and (10.20), DN = L1 (S1 R1 L2 S2 )R2 , the Smith form of
DN is the same as that of S1 R1 L2 S2 because L1 and R2 are unimodular. It
is therefore sufficient to prove that the Smith form of S1 M S2 is S1 S2 , where
10.1 General Disturbances: Internal Model Principle 337
= (d1k d−1 −1
1,k−1 )(d2k d2,k−1 )
= λ1k · λ2k ,
which implies that the Smith form of Q = S1 M S2 is S1 S2 , The proof is thus
complete.
Theorem 10.1.2. Let a full row rank N ∈ Rm×l [s] and a nonsingular
m×m
D∈R [s] be given with σ(D) ∩ σ(N) = ∅. Then N and D are externally
skew prime and thus externally skew prime in C+ as well. Furthermore, D̄
and N̄ , defined by
S1 0
D̄ = R3 , (10.26)
0 I
N̄ = L3 S2 , (10.27)
338 10. Asymptotic Disturbance Decoupling
DN = L3 S3 R3 = L3 S1 S2 R3
= L3 [Λ1 Λ2 0]R3
Λ1 0
= L3 [Λ2 0] R3
0 I
S1 0
= L 3 S2 R3
0 I
:= N̄ D̄,
We construct
10.1 General Disturbances: Internal Model Principle 339
+
0 0 0 0
+
G̃d: = (Gd Ḡd ) = 0.2 0.8
= 0.2 0.8
.
s(2s+1) s(2s+1) s s
Consider first the case of SISO linear time-invariant (LTI) continuous pro-
cesses. Let G(s) be the plant and K(s) the feedback controller. The proposed
scheme, called the virtual feedforward control scheme (VFC scheme for short),
is depicted in Figure 10.4, where d is supposed to be an unknown periodic
disturbance. Without the VFC controller, the proposed structure reduces to
an ordinary feedback control system and in face of non-zero periodic d, the
system will have non-zero steady state error. The intention of the proposed
method is to activate the VFC controller timely and to give an extra control
10.2 Periodic Disturbances: Virtual Feedforward Control 341
It is well known that the steady state of the output y(t) of a stable system
G in response to a periodic input x(t) is periodic with the same period as
that of the input. With the standard convolution operation ∗, the response
of a stable system Y = G2 G1 X can be written as y(t) = g2 (t) ∗ g1 (t) ∗ x(t),
where g1 (t) and g2 (t) are impulse responses of G1 and G2 , respectively. For
simplicity, denote g2 (t) ∗ g1 (t) ∗ x(t) by (g2 ∗ g1 ∗ x)(t). For a signal f (t) with
a periodic steady state, let its steady state be f s (t), and the corresponding
Laplace transform be F s (s) = L{f s (t)}. Lemma 9.3 in Chapter 9 is needed in
our development here and repeated as follows (readers are referred to Chapter
9 for its proof).
Lemma 10.2.1. Under a periodic input x(t), the output steady state of the
system Y = G2 G1 X with stable G2 and G1 satisfies
With this lemma in hand, we are now ready to develop the new method
for periodic disturbance rejection. Suppose that the process is in a normal
and stable feedback control (v = 0 in Figure 10.4) and the output has been in
a constant steady state before the disturbance acts on the system. Without
loss of generality, such a steady state is assumed to be zero. When a periodic
disturbance d comes to the system at t = 0, the process output y will become
342 10. Asymptotic Disturbance Decoupling
periodic with the same period after some transient. The periodic output can
be detected and its period be measured with high accuracy by monitoring
the process output.
At t = 0, the VFC controller has not been activated yet (v = 0 in Fig-
ure 10.4), and it follows from Figure 10.4 that Yd = D + GUd , or
D = Yd − GUd , (10.28)
where the subscription d indicates the signals’ response to d only. Note that
yd (t) and ud (t) as well as their steady states yds (t) and usd (t) are available. In
the time domain, (10.28) becomes
Since the actual process transfer function G is not available, its model Ĝ has
to be used to estimate d(t), i.e.,
ˆ = yd (t) − (ĝ ∗ ud )(t),
d(t) (10.30)
or in the s-domain,
D̂ = Yd − ĜUd . (10.31)
Note that d(t) and d(t)ˆ have the periodic steady state ds (t) and dˆs (t) respec-
tively, since yd (t) and ud (t) are so and the system is stable. Now suppose that
the VFC controller v(t) is activated at t = Tv . Since the initial conditions of
the system at t = Tv are non-zero, we have to apply the Laplace transform
at t = 0 to get
1 G
Yv = D− V, (10.32)
1 + GK 1 + GK
where the subscription v indicates the signal’s response to both d and v. Let
1(t) be the unit step function. We set the virtual feedforward control signal
as
Let
1
Qvd = Ĝ−1 (s)D̂(s), Hyd = ,
1 + GK
Equation (10.33) can be written in the time domain as
Choose
Tv = kT, (10.36)
where T is the period of d(t), yds (t) and usd (t), and k is an integer, so that
there hold
which becomes
1 − GĜ−1
F ,
1 + GK
and f as its impulse response. (10.38) is written as
Introduce the following norm for a periodic signal x(t) with the period of T :
T
x(t) = x2 dt,
0
T ∞
yds (t)2 = [yd0 + Ci cos(ωi t + φi )]2 dt. (10.40)
0 i=1
Noting the orthogonality of cos(ωi t) and cos(ωi t+φi ) on the interval t ∈ [0, T ],
(10.40) gives
∞
T
yds (t)2 = T yd0
2
+ Ci2 . (10.41)
2 i=1
Similarly, we have
T ∞
yvs (t)2 = [F (0)yd0 + Ci |F (jωi )| cos(ωi tφi + ∠F (jωi ))]2 dt
0 i=1
∞
T
= F 2 (0)T yd0
2
+ Ci2 |F (jωi )|2 . (10.42)
2 i=1
If
|F (jωi )| < 1, i = 1, 2, · · · , (10.43)
it follows from (10.42) that
yvs (t) < yds (t). (10.44)
In the VFC scheme, the extra control v is activated according to (10.34),
(10.30) and (10.36). This v is independent of the closed-loop system once
activated and acts as a “feedforward” control with the estimated disturbance
ˆ Therefore, stability of the closed-loop system can be guaranteed with the
d.
proposed VFC strategy, provided that the stability is taken into consideration
at the stage of controller design. In view of the above development, we obtain
the following theorem:
Theorem 10.2.1. A stable feedback system remains stable when the virtual
feedforward control signal v described by (10.34), (10.30) and (10.36) is in-
troduced, and the resultant output steady state in response to a periodic dis-
turbance satisfies
yvs (t) = 0, if Ĝ = G,
−1
yvs (t) < yds (t), if | 1−G(jω i )Ĝ(jωi )
1+G(jωi )K(jωi ) | < 1, i = 1, 2, · · · .
behavior is observed, the VFC controller is activated: the steady state of the
disturbance d is estimated from (10.30) and v is computed from (10.34) and
applied to the system at Tv according to (10.36).
It can be seen from (10.45) that once the transient xt (t) dies out, the steady
state part xs (t) can be found by
xs (t) = [g −1 (t) ∗ y(t) − βeat ]s . (10.46)
The extension of the method to non-minimum-phase systems with multiple
RHP zeros is straightforward. In implementation, some numerical integration
technique (Sinha and Rao, 1991) has to be used to compute β, and this
inevitably gives a numerical error, which will accumulate and make η(t)
g −1 (t) ∗ y(t) − βeat diverge. Thus, we only take one early period of η(t), when
η(t) becomes approximately periodic, and duplicate it by η s (t) = η s (t + T )
to generate xs (t) = η s (t) for all t 0.
For example, consider a non-minimum-phase process G(s) = −4s+1 5s+1 , and
1
suppose that a sinusoidal signal x(t) passes through 9s2 +4s+1 and then acts
on the process. With the suggested method, the steady state periodic signal
x̂s (t) is constructed and shown in Figure 10.5, where the solid line is x̂s (t)
and the dashed line is the actual x(t). It is clear that x̂s (t) is almost identical
to the steady state part of x(t).
0.25
0.2
0.15
0.1
0.05
−0.05
−0.1
−0.15
0 5 10 15 20 25 30 35 40 45 50
−4s+1
Fig. 10.5. Non-minimum-phase System G(s) = 5s+1
(—— constructed x̂s (t), − − − actual x(t))
Thus, (10.47) and (10.48) may be called static and dynamic VFC, respec-
tively. One may guess that the dynamic control (10.48) can improve distur-
bance rejection performance over the static one (10.47). Dynamic VFC and
static one have the same steady state, which causes the same steady state
output response. The difference between the signals generated by these two
schemes is its transient which depends on the plant and disturbance which
vary from the case to the case, and is hard to predict. This difference will
cause the transient difference in the output response and the latter is even
harder to predict. Thus, there is no guarantee that the dynamic VFC will
give an improved (dynamic) performance over the static VFC. What we see
from extensive simulation is that for some examples, dynamic VFC does
give a little improvement over static VFC, while for most examples, almost
no difference can be observed. Furthermore, the dynamic VFC is not appli-
cable to non-minimum-phase models. Therefore, we recommend (10.34) for
practical implementation for all stable models of both minimum-phase and
non-minimum-phase.
a new steady state with a suppressed error is expected after some transient.
Thereafter, one waits for a different output waveform to occur, and then the
VFC controller is de-activated (or removed) from the system. The procedure
is then repeated.
We now present some different scenarios to demonstrate the effectiveness
of the VFC scheme.
Example 10.2.1. Consider a process:
e−4s
G(s) =
(10s + 1)(2s + 1)
under a PID controller (Zhuang and Atherton, 1993):
0.2631
K(s) = 2.0260 + + 4.9727s. (10.49)
s
The external disturbance signal is introduced at t = 0:
2
1
d
0
−1
2
1
y
0
−1
0 50 100 150 200 250 300 350 400
5
0
u
−5
4
2
v
0
−2
1.03e−5.47s
Ĝ(s) = .
11.41s + 1
From the recorded y(t), v(t) is computed with Ĝ(s) and applied with Tv =
3T ≈ 188.5. The performance is shown in Figure 10.8 over t ∈ [0, 377.0],
where the solid line is with model mismatch, while the dashed line is for the
case without model mismatch.
10.2 Periodic Disturbances: Virtual Feedforward Control 349
0
y
−1
−2
−3
1.5
1
0.5
0
d
−0.5
−1
−1.5
0
y
−1
−2
−3
0 100 200 300 400 500 600 700 800
Such a disturbance will change the waveform of the monitored y(t), and
a steady state periodic y(t) with enlarged error is observed. The adaptive
scheme is thus activated to de-activate v(t) at t = 7T ≈ 439.8. The virtual
feedforward block (with model mismatch) is activated again at t = 10T ≈
628.3 to compensate for the new disturbance, as shown in the solid line in
Figure 10.8. The response without the adaptive scheme, i.e., with the same
v(t) for the different d(t), is shown by the dashed line over t ∈ [377.0, 800.0]
in Figure 10.8.
In practice, measurement noise is generally present, and its level may be
measured by the noise-to-signal ratio (Haykin, 1989):
mean(abs(noise))
N SR = .
mean(abs(signal))
Example 10.2.2. The proposed method can be applied to servo control prob-
lems for better disturbance rejection over the existing methods, say the repet-
itive control scheme. For demonstration, consider a servo system in Moon et
al. (1998) with the uncertain plant:
3.2 × 107 β0
G(s) = ,
s2 + α1 s + α0
where α1 ∈ [20, 27], α0 ∈ [45000, 69000], β0 ∈ [60, 100]; and the controller:
We compare our method with the repetitive control scheme discussed in Moon
et al. (1998). Suppose that the system is in the steady-state before a distur-
bance is introduced at t = 0, the disturbance signal comprises a constant
signal, eight out-of-phase sinusoidal signals from the first to the eighth har-
monics of the spindle frequency, 30 Hz (Moon et al., 1998).
The coefficients of the model are supposed to be center values of each
interval, i.e.,
1.5
0.5
0
d
−0.5
−1
−1.5
(a) disturbance
0
y
−1
−2
−3
0
y
−1
−2
−3
For the case without model mismatch, i.e., the actual process is described
exactly by the model, the responses are shown in Figure 10.10, where the
solid line is from the proposed method, and the dashed line is from the
repetitive control scheme. Since the control signals from these two methods
are very close to each other, their difference ∆u is displayed in Figure 10.10,
352 10. Asymptotic Disturbance Decoupling
50
d (um)
−50
0.5 0.51 0.52 0.53 0.54 0.55 0.56 0.57 0.58 0.59
0.02
0.01
y (um)
−0.01
−0.02
0.5 0.51 0.52 0.53 0.54 0.55 0.56 0.57 0.58 0.59
0.05
0
u
−0.05
0.5 0.51 0.52 0.53 0.54 0.55 0.56 0.57 0.58 0.59
−4
x 10
3
1
∆u
−1
−2
0.5 0.51 0.52 0.53 0.54 0.55 0.56 0.57 0.58 0.59
Fig. 10.10. Comparison of VFC and Repetitive Control (—— VFC, − − − Repet-
itive)
where only the steady state parts are shown for illustration. The effectiveness
of the proposed method is obvious. Now we vary the plant coefficients within
the possible intervals:
The responses are shown in Figure 10.11. The robustness of the proposed
method to the model uncertainty is noticed. ♦
10.2 Periodic Disturbances: Virtual Feedforward Control 353
50
d (um)
−50
0.5 0.51 0.52 0.53 0.54 0.55 0.56 0.57 0.58 0.59
0.04
0.02
y (um)
−0.02
−0.04
0.5 0.51 0.52 0.53 0.54 0.55 0.56 0.57 0.58 0.59
0.05
0
u
−0.05
0.5 0.51 0.52 0.53 0.54 0.55 0.56 0.57 0.58 0.59
−4
x 10
6
2
∆u
−2
−4
0.5 0.51 0.52 0.53 0.54 0.55 0.56 0.57 0.58 0.59
We have so far considered the SISO case. The extension to the MIMO case
is straightforward. Suppose that the process P (s) is invertible and has been
stabilized by the controller K in a conventional feedback system. The same
scheme as in Figure 10.4 is exploited. The same derivation and implementa-
tion techniques as in the SISO case before applies with obvious modifications
to suit the square matrix case. The only thing to take note is that the inverse
354 10. Asymptotic Disturbance Decoupling
of the process G is needed in (10.34) to compute the VFC vector signal v(t)
and is usually highly complicated and not realizable. We can apply the model
reduction (Chapter 2) to G−1 (jω) to get its approximation G̃(s) in form of
g̃ij (s) = g̃ij0 (s)e−τij s . (10.52)
−τij s
where g̃ij0 (s) are proper rational functions. Some terms e may happen to
be a pure predictor but by choosing Tv = kT with a suitable k (see (10.36)),
we can always make all τij negative, leading to time delays terms e−τij s which
are realizable.
Example 10.2.3. Consider the process:
12.8 −s −18.9 −3s
16.7s+1 e 21.0s+1 e
G= 6.6 −7s −19.4 −3s
,
10.9s+1 e 14.4s+1 e
which is controlled by
0.949 + 0.1173/s + 0.4289s 0
K= .
0 −0.131 − 0.0145/s − 0.1978s
With model reduction, we obtain
G−1 ≈ G̃(s)
0.31s +0.62s3 +0.42s2 +0.05s+0.002 2s
4
0.089s4 +0.04s3 −0.19s2 −0.03s−0.0015 4s
s4 +0.25s3 +1.16s2 +0.21s+0.013 e s4 +0.36s3 +1.03s2 +0.17s+0.01 e
= .
0.14s4 +0.278s3 +0.18s2 +0.02s+0.0007 −2s 0.08s4 +0.03s3 −0.16s2 −0.03s−0.001 6s
s4 +0.26s3 +1.16s2 +0.22s+0.01 e s4 +0.38s3 +1.03s2 +0.19s+0.01 e
0.5
0
d
−0.5
−1
0.2
0.15
0.1
1
v
0.05
−0.05
0.1
0.05
−0.05
2
v
−0.1
−0.15
−0.2
1.4
1.2
0.8
y1
0.6
0.4
0.2
0
50 100 150 200 250 300 350 400 450 500
1.5
y2
0.5
0
50 100 150 200 250 300 350 400 450 500