Introduction to Mathematical Control Theory
Introduction to Mathematical Control Theory
Introduction
~to Mathematical
-Control Theory
- STEPHENBARNETT
| a De es
~~,
Introduction to
mathematical
control theory
ISBN 0 19 859619 7 9
©OxfordUniversityPress1975
All rights reserved. No part of this publication may be reproduced,
Stored in a retrieval system, or transmitted, in anyform or by any means,
electronic, mechanical, photocopying, recording or otherwise, without
the prior permission af Oxford University Press
Control theory has developed rapidly over the past two decades, and is
now established as an important area of contemporary applied mathe-
matics. Some justification for this latter claim is provided in Chapter 1,
which begins with a general discussion on the scope and relevance of
control theory. The aim of this book is to present a concise, readable
account of some basic mathematical aspects of control. The origins of
the book lie in courses of lectures which I have given for a number of
years to final honours mathematics students and to postgraduate control
engineers at the University of Bradford. My approach concentrates on
so-called state space methods, and emphasises points of mathematical
interest, but I have avoided undue abstraction and the contents should
be accessible to anyone with a basic background in calculus and matrix
theory. To help achieve this objective a summary is given in Chapter 2
_of some appropriate topics in linear algebra. The book should therefore
also be useful to qualified engineers seeking an introduction to con-
temporary control theory literature, some of which tends to be expressed
with rather forbidding mathematical formulation.
I have provided problems to be attempted at the end of each section.
Most of these have been class-tested, and so should be within the reader’s
grasp; all answers are given. In an introductory work it is essential to
supply references for further study, and except in a few cases I have
quoted text books rather than journal papers which are usually more
suitable for specialists. It is of interest that most of the references listed
have appeared within the last ten years, giving some measure of the
publishing ‘explosion’ in the control field in recent times.
In any compact treatment many points of detail, refinements and
extensions are inevitably omitted; in particular I regret that it has not
been possible to include any material on statistical concepts which arise
in control. Despite the many texts now available on control theory it is
my hope that this book fills a niche in the literature by its combination
of conciseness and range of topics. I hope also that it will help to pro-
mote interest in this exciting field, thus providing at least a small counter-
balance to the deadweight of British classical applied mathematics.
I havedonemy bestto eliminateerrors,althoughin an entertaining
vi Preface
article P. J.Davis (Amr. Math. Mon., 1972, 79, 252) suggests that such
attempts are doomed to failure! Thanks however are due to P. A. Cook,
A. T. Fuller, J. B. Helliwell, G. T. Joyce, A. G. J. MacFarlane, D. D. Siljak,
C. Storey, and H. K. Wimmer who have made useful comments on various
parts of the manuscript. I should also like to thank Dr. Hedley Martin for
encouraging me to write this book and for his helpful assistance through-
out the project, and the Clarendon Press for their efficient co-operation.
Finally, as on many previous occasions, I am grateful to Mrs. Margaret
Balmforth for her skilful preparation of the typescript.
Bradford S.B.
July 1974
Contents
See OLABIEITY
5.1. Definitions
5.2. Algebraic criteria for linear systems
5.2.1. Continuous-time
Annn
DPW Liapunov theory
Application
Construction
5.6.1.
of Liapunov theory to linear systems
of Liapunov functions
Variable gradient method
5.6.2 Zubov’s method
5.7. Stability and control
5.7.1. Input—output stability
5.7.2. Linear feedback
5.7.3.+ Nonlinear feedback
6. OPTIMAL CONTROL
6.1. Performance indices
6.1.1. Measures of performance
6.1.2. Evaluation of quadratic indices
6.2. Calculus of variations
6.3. Pontryagin’s principle
6.4. Linear regulator
6.5. Dynamic programming
REFERENCES
ANSWERS TO EXERCISES
INDEX 261
Disturbances
Control
variables
(inputs) Controlled
system Outputs yi,
Controlling:
[5175
Deen)
oy state
variables
Sy i=) Te
devicecnet :
oeFes
|VP) Output
monitoring
Feedback
FIG/1.1
and any input then the resulting state and output at some specified later
time are uniquely determined. Such models are often termed dynamical
systems, although of course they need have nothing to do with Newtonian
mechanics, and can be defined in rigorous formal terms (e.g. Desoer 1970).
If the controller operates according to some pre-set pattern without
taking account of the output or state, the system is called open loop,
because the ‘loop’ in Fig. 1.1 is not completed. If however there is feed-
back of information concerning the outputs to the controller, which then
appropriately modifies its course of action, the system is closed loop.
Simple illustrations of open and closed loop systems are provided by
traffic lights which change at fixed intervals of time, and those which are
controlled by some device which measures traffic flow and reacts accord-
ingly.
Another familiar example is provided by the closed loop feedback
system which controls the water level in the cistern of the domestic toilet.
After flushing is completed the water outlet valve closes, and water flows
into the cistern. A ball float measures the water level, and closes the
supply valve when the water reaches a pre-set level. Thus a knowledge of
the actual water level is ‘fed back’ via the float to obtain a measured level
which is compared with the desired level; when this difference, or ‘error’
e, exceeds a predetermined value the supply valve is open, and when e is
less than this value the valve is closed. This is represented diagrammatically
in Fig. 1.2.
FIG. 1.2 @= actual water level; d = desired water level; m = measured water level; e = d—m.
Both Figs 1.1 and 1.2 are examples of block diagrams, which are widely
used in control engineering as a convenient method of representing the
relationships within a system in a pictorial form. The convention used is
that flows are added with their indicated signs where the junction is shown
as a circle, otherwise signals proceed along branching paths without altera-
tion. The rectangles represent operators of some sort.
4 Introduction to control theory
In the plumbing system it is clear that the control of water level will
be satisfactory even if there are wide variations in the pressure of the in-
flowing water. Similarly, in the central heating system mentioned earlier,
the performance may well be acceptable even if the heat output of the
boiler varies or if outside weather conditions change suddenly. These
examples illustrate a very important property of feedback systems: their
ability to operate satisfactorily even when parameters in various parts of
a system may vary considerably. This insensitivity will be discussed in a
little more detail in Section 1.2. Feedback is thus a crucial concept in
control engineering, and indeed a water level control of the type described
above was known to the Greeks in the first century A.D. Although feed-
back has a long history (see Mayr 1970), it is only in the last forty years
or so that extensive efforts have been given to the study of feedback
theory. The so-called ‘classical’ control engineering methods rely heavily
on transform methods for solving linear differential or difference equa-
tions, and received great impetus from military applications during the
Second World War. Before going into further details in the next section,
we first discuss some aspects of ‘modern’ control theory, which has been
developed mainly over the last twenty years using the state variable
approach, and on which this book concentrates most attention. The
examples presented below will be referred to again in subsequent chap-
ters, and all illustrate the important feature of control theory, mentioned
earlier, namely to synthesize a control strategy which satisfies require-
ments.
Example 1.1. Suppose a car is to be driven along a straight, level road,
and let its distance from an initial point 0 be s(t) at time ¢. For simplicity
assume that the car is controlled only by the throttle, producing an acce-
lerating force of u,(t) per unit mass, and by the brake which produces
a retarding force of u(t) per unit mass. Suppose that the only factors of
interest are the car’s position x, (f) = s(t) and velocity x, (t) = s(t), where
(*) =d/dt. Ignoring other forces such as road friction, wind resistance,
etc., the equations which describe the state of the car at time f are
X1 =X2
Xq =U, — U2
or in matrix notation
%=Ax + Bu (1.1)
Introduction to control theory 5
where
X41
| head uy 01
Placa (allPg eeden P= 00
xX Uy 00 Pt
In practice there will be limits on the values of u, and u,, for obvious
reasons, and it will also be necessary to impose restrictions on the magni-
tudes of velocity and acceleration to ensure passenger comfort, safety,
tc;
It may then be required to start from rest at 0 and reach some fixed
point in the least possible time, or perhaps with minimum consumption
of fuel. The mathematical problems are firstly to determine whether such
objectives are achievable with the selected control variables, and if so, to
find appropriate expressions for uv, and uw,as functions of time and/or
x, and x2.
The complexity of the model could be increased so as to take into
account factors such as engine speed and temperature, vehicle interior
temperature, and so on. A further aspect of realism could be added by
imagining the car to be travelling in one lane of a motorway, in which
case the objective might be to maintain the distance from the vehicle in
front within certain limits.
ot
it
Heating
coil
FIG.-le3
6 Introduction to control theory
Let the heat capacities of the oven interior and of the jacket be c,
and c, respectively; let the interior and exterior jacket surface areas be
a, and a; and let the radiation coefficients of the interior and exterior
jacket surfaces ber; andr. Assume that there is uniform and instan-
taneous distribution of temperature throughout, and that rate of loss
of heat is proportional to area and the excess of temperature over that
of the surroundings. Ignoring other effects, if the external temperature
is J), the jacket temperature is 7, and the oven interior temperature is
T>, then we have:
Forthejacket:
eyT;=—ayro(T,
—To)—a17,
(TyTa)ee (1.2)
Fortheoven
interior:
caT2=440;
aged): (1.3)
Let the state variables be the excesses of temperature over the exterior,
ie.x, =T, —To,X2 =T2 —To. It is easy to verify that (1.2) and (1.3)
can be written in the form (1.1) with
lytically and numerically, and the first model of a situation is often con-
structed to be linear for this reason. Furthermore, the technique of
linearization of nonlinear systems is an important one, and relies on the
fact that if the perturbation z(t) from some desired state x(t) is small then
a set of linear equations in z can be formed by neglecting all but the first
terms in a Taylor series expansion in z. Chapters 3 and 4 are devoted to
the study of basic properties of linear systems. Of course in many cases
linear descriptions may be inapplicable, as the following example illustrates.
FIG. 1.4
The state variables are the velocity x, of S, relative to S,, the distance
Xz between S, and S, and the angle x3. Ignoring any external gravitational
forces, the state equations are
x, =ccos(u+x3)
X2 =—X; cos x3
The problem might then be to find a function wu,if possible, which takes
S; to S, in minimum time T or with minimum consumption of fuel. A
numerical solution to the former problem, with x; ae 0, can be found
in Bryson and Ho (1969, p. 143).
8 Introduction to control theory
FIG.1.5
me =(Y
sin 0 —X cos 0)
d2
X=m sn (z + sin 6)
a (1.5)
Y—mg=mon (2 cos8)
u X=M
melea2»
where X and Y are the horizontal and vertical components of the force
exerted by the hinge on the rod, and z is the displacement of the centre
of mass of the platform from some fixed point. Again, an initial problem
is to determine whether the desired objective can be achieved.
Introduction to control theory 9
Example 1.5. Suppose that the sales S(¢) of a product are affected by
the amount of advertising A(t) in such a way that the rate of change of
sales decreases by an amount proportional to sales, but increases by an
amount proportional to the advertising applied to the share of the market
not already purchasing the product. If the total extent of the market is
M, the state equation is therefore
S =-aS + bA(#)(1 -S/M)
Exercise 1.1.
In Example 1.1, list as many factors as you can think of which are present
in reality and can affect the motion of a car being driven along a road.
Exercise 1.2.
Consider the system composed of two masses lying on a smooth hori-
zontal table and connected by two springs to a fixed support, as shown
below.
10 Introduction to control theory
YASpring Spring
constant constant
2
Exercise 1.3.
In the inverted pendulum problem in Example 1.4 take as state variables
x, =0,x, =0,x3 =Z,x4 =Z, and set u = 0. By eliminating X and Y in
eqns (1.5), obtain the system equations in state space form X = f(x, x).
Exercise 1.4.
Consider a controlled environment consisting of rabbits and foxes, the
numbers of each at time ¢ being x, (t) and x2 (ft) respectively. Suppose
that without the presence of foxes the number of rabbits would grow
exponentially, but that the rate of growth of rabbit population is reduced
by an amount proportional to the number of foxes. Furthermore suppose
that, without rabbits to eat, the fox population would decrease exponen-
tially, but that the rate of growth in the number of foxes is increased by
an amount proportional to the number of rabbits present. Show that
under these assumptions the system equations can be written
Xy =X, —A2X%2,X2q
= 3X1 —dgXx2 (1.6)
where the a; are positive constants.
This system will be studied again in Exercises 1.7, 4.15, and 4.33, and
Example 5.2.
Introduction to control theory Nia
Classical control theory deals with a linear time invariant systems having
scalar input u(t) and scalar output z(t). The analysis problem is to study
the behaviour of a given system in specified circumstances. The aim of
design is to ensure that the performance of the system satisfies required
specifications. If z(¢) is required to be as close as possible (in some sense)
to a given reference signal r(t) the control system is called a servomecha-
nism, and if r is constant, a regulator. For example, the inverted pendulum
discussed in Example 1.4 is a regulator, since it is required to keep the
angle to the vertical 6 as near to zero as possible. Similarly, a central-
heating system is designed to keep room temperature close to a prede-
termined value.
Z(s)=L{
z(t)}=
its2(t)e~*
dt. (1.9)
It is assumed that the reader has a working knowledge of the use of Lap-
lace transforms for solving ordinary differential equations.
Assuming that all the 2(0) and u(0) are zero, application of (1.8)
to (1.7) gives
k(s) 2(s) = B(s)u(s) (1.10)
12 Introduction to control theory
where
K(s)=s" +k, s® 1 400 e4hk, 54K, (1.11)
PIGS 1.7
If u(t) (and hence u(s)) is known then solution for z(t) requires expan-
sion of the right hand side of eqn (1.13) into partial fractions. If the
poles ofg(s) are denoted by A,, A2,..., Ay (these of course may be real
or complex numbers) then this expansion will involve terms of the form
c;/(s —Ay),the c; being constants. These correspond to terms cjexp(Ajt)
in z(t). Solution of linear systems is dealt with systematically in Chapter
3 using matrix methods, which are generally preferable unless n is small.
It is worth noting that if the system is unforced (i.e. u= 0) and is subject
to given initial conditions at ¢ = 0, then the solution for z(f) (i.e. the com-
plementary function for (1.7)) tends to a steady state ast > © only if
all the A; have non-positive real parts. Constant terms in the steady state
solution correspond to zero As, and sinusoidal terms to purely imaginary
As. The question of stability of linear systems is gone into in detail in
Chapter 5.
Classical control theory is based on the study of transfer functions,
and a number of powerful methods have been evolved. However, with
one exception (see Section 5.3) these will be dealt with only very briefly
in this book. This is not to deny the enormous practical utility of the
techniques, but reflects the fact that for the mathematician, at least,
Introduction to control theory 13
greater interest lies in the more recent approaches to the study of control
systems, especially those involving vector inputs (multivariable control
systems). Moreover, transform methods are applicable only to constant
linear systems and are of no help with time varying or nonlinear systems.
Consider a closed loop system represented in the block diagram below
Fig. 1.8, where g(s) is the open loop transfer function and A(s) is the
feedback transfer function.
FIG. 1.8
4 se)
&(s) T+esi)’ (1.15)
u(t)=1,t>0
(1.16)
=0,t<0
and that
IT @~ TT 6-4)
i=1 Jeu
14 Introductionto controltheory
where all the poles 41, ..., My of the closed loop system transfer function
are assumed distinct. Then it is easy to show using u(s) = 1/s that the
system response to the step input is
We have seen that system response depends on the poles and zeros of
g-(s). Suppose that the open loop transfer function g(s) is multiplied by
a gain parameter K so that (1.15) becomes
8c (8) =Kg(s)/[1 + Kg(s)r(s)]. (1.21)
The root locus method has been devised for investigating how the poles
of g_ (s) in (1.21) vary with K. These poles are the roots of
1 + Kg(s)h(s) = 0 (1.22)
and sets of rules have been drawn up for determining the loci of these
roots graphically as K varies from 0 to -. Again, the mathematical detail
is rather tedious.
There are very many textbooks available on control engineering and
the reader should consult these for comprehensive accounts of the above
and other classical methods. In particular the books by Elgerd (1967),
Eveleigh (1972) and Takahashi, Rabins, and Auslander (1970) can be
recommended for treatments which combine the classical and modern
approaches, and the account by Truxal (1972) is clear and interesting.
It is instructive to return in more detail to the question of sensitivity
of feedback systems, briefly mentioned in Section 1.1. For a transfer
function ¢(s) the sensitivity with respect to small variations in some
parameter 0 is defined by
deft _ dt.
Se) ~a6/6 a0 Ps (1.23)
Consider the systems represented in Fig. 1.9, and
16 Introduction to control theory
suppose for simplicity that g,,g2, and h are constants. For the open loop
system in Fig. 1.9(a) the transfer function is just t; =18 so Sg (t,) = 1.
Thus a small percentage change in g, causes the same percentage change
in t,. For the closed loop system in Fig. 1.9(b) eqn (1.15) gives
Clearly in this case for a given gz, the values of g; and h can be chosen so
as to make the sensitivity to small changes in g2 as small as we please. Such
sensitivity considerations carry over in a fairly general way to more com-
plicated systems, and thus provide a very important practical justification
for the use of feedback. In a similar way it can be demonstrated that feed-
back can reduce the effect of external disturbances on a system output.
As is illustrated by (1.25), the denominator of the transfer function plays
an important role in these discussions. More generally, the denominator
1 + g(s)h(s) in the transfer function (1.15) is termed the return difference
for the following reason. Suppose that in the corresponding block diagram
(Fig. 1.8) uw = 0, but that an input e = 1 is inserted. The returning signal
will be —g(s)h(s), so the difference between them is just 1 + gh.
We shall encounter the ideas of transfer functions in our subsequent
work on linear systems, especially the extension (for multi-input multi-
output systems) to matrices whose elements are transfer functions.
We now turn to situations where the variables are measured (or ‘sampled’)
only at discrete intervals of time instead of continuously, producing what
are referred to as sampled-data, or discrete-time systems. Such circum-
stances are common in real life: for example, the interest on savings
accounts is calculated daily or monthly; the temperature of a hospital
patient is recorded at perhaps hourly intervals; a driver of a car glances
at the instruments, rear mirror, etc., only intermittently —the reader can
easily add many instances of his own (see Cadzow 1973, ch. 2).
We shall suppose that our variables are defined at fixed intervals of
time 0, T, 27, 37, . . . where T is a constant, and use X(k), u(k) to denote
the values of the output X (KT), and the input u(kT) respectively (k = 0,
Introduction to control theory 17
X@)
=2{xo}
=>,xG02* (1.27)
k=0
=x)+707D. . (1.28)
The reader may not have encountered z-transforms previously so it will
be useful to go into a little detail here.
=z/(z-1).
(b) Similarly
2
dt a,a tat.
Blab =lt+—+steee
a4
=(lire)
= z/(z —a). (1.29)
18 Introduction to control theory
Z{X(k
+) }=z!X(z)-z/X(0)—-z!~
1y(1)-z/-?x(2)- 20—z*XG=2)-
zX(j-1) (131)
which is the analogue of the result for the Laplace transform of the jth
derivative of z(t) given in (1.8). Application of (1.31) to the difference
equation (1.26) then transforms this into an algebraic equation for X(z).
The resulting expression is expanded into partial fractions as for Laplace
transforms.
=2ec +X1) , GX 5)
zi £42
Hence using (1.29) the solution of the equation is
X(k)=—(2X9+X1) —3)*+3X0 +X1)C2).
Introduction to control theory 19
X(z) =g(z)u(z)
where g is defined by (1.14). Some of the continuous-time frequency
domain methods have been extended to deal with sampled-data transfer
functions (see for example Saucedo and Schiring 1968).
Exercise 1.5. ;
' Show that the system represented in Fig. 1.8. is equivalent to that shown
in Fig. 1.10.
g(s)A(s)
FIG, 1.10
Exercise 1.6.
Find the transfer function for the system represented by the block dia-
gram in Fig. 1.11.
Exercise 1.7.
Consider the rabbit-fox environment described in Exercise 1.4. If the
experiment begins with initial population sizes x, (0) and x2(0) and if
20 Introduction to control theory
FIG. 1.11
in eqns (1.6) a,/a3 =a2/aq, use Laplace transforms to solve for x; (f).
Hence show that for arbitrary initial conditions the populations will
attain a steady state ast > c only if a; —aq4<0, and give an expression
for the ultimate size of the rabbit population in this case. Finally, deduce
that if the environment is to reach a steady state in which both rabbits
and foxes are present then x, (0) > (a, /a3 )x2(0).
What happens if x, (0) = (a, /a; )x.(0)?
Exercise 1.8.
Deduce by considering eqn (1.22) that the root loci start (i.e. K= 0) and
terminate (i.e. K = °°) at the poles and zeros respectively of the function
8(s)h(s).
Exercise 1.9.
For the transfer function obtained in Exercise 1.6. derive an expression
for the sensitivity with respect to g,. How can this be made equal to zero?
Exercise 1.10.
Obtain the z-transforms of the following functions, defined for k = 0, 1,
2,... (a) X(k) = kT (b) X(k) = e%, where a is a constant.
By replacing a by ia in (b) and considering imaginary parts, deduce that
{ d } z sina
z*—-2zcosatl.
Zysinak ¢=————_
Exercise 1.11.
Prove that
Z{e?*X(k)}=
X(ze)
Introduction to control theory 21
Exercise 1.12.
A very simple model of a national economy can be constructed by assum-
ing that at year k the national income J; is given by
Te=2{1+(1/4/2)*sin %k
IG
(the result in (1.32) is needed).
The assumptions of the model therefore lead to a national income
which is oscillatory, and as k > © national income tends to twice govern-
ment expenditure.
Exercise 1.13.
The following simplified model for the buffalo population in the American
West in 1830 has been suggested by Truxal (1972). Let Fx, Mybe the
numbers of female and male buffalo at the start of year k (k = 0 corres-
ponding to 1830). Five per cent of adults die each year. Buffalo reach
maturity at age two years, and the number of new adult females alive
at the beginning of year k + 2, taking into account infant mortality, is
22 Introduction to control theory
12 per cent of Fx; more male calves than female are born, and the corres-
ponding figure is 14 per cent of Fz. Thus
Fy 49 ="0:95F,., + 0-12F, CY.33)
Use z-transforms to find expressions for F; and Mx, and deduce that as
k > © the population increases by 6.3 per cent per year.
This model will be studied further in Exercise 4.53 and Example 5.11.
2 Preliminary matrix theory
for the unit matrix of order n which has the property that Al, =/,A for
any n X n matrix A. The matrix in (2.2) has all its elements zero except
24 Preliminary matrix theory
those on the principal diagonal, and any square matrix of this form is
called a diagonal matrix, written diag [a11,4@22,...,4nn]. It is assumed
that the reader is familiar with the usual rules of addition, subtraction,
multiplication, inversion, etc. for matrices. If the elements in any kK(<m)
rows and &(<n) columns of A are selected they form ak x 2 submatrix of
A, In particular A can be partitioned into submatrices (or blocks) aj; of
orders mj; x nj, as follows:
ny Na ° ° Ng
my, mM . My
O11 | 0 ; 0 my,
eS ee ee Ler
| 1 |
c iss
diag aes. «ss 0 |1 O22| Neer | 0 mM 3
at Sl ee es 4--- (2.3)
eu,
rr " 1 !
e I e 1 e | e °
| I |
across howe (rae
O- wsiQonk T *? hlebes Mr
where
x =([X11,X12,
s203XimsX215+++sX2amo->>
xa | (2.8)
is the column mn-vector formed from the rows of X taken in order, and
c is formed in an identical fashion from C. Similarly, XB = C can be
written
if A is diagonal or triangular, i.e. all the elements above (or below) the
principal diagonal are zero, so that a;,= Oif7 >i (orj <i). If det A =0,
A is singular, otherwise nonsingular; in the latter case the inverse of A is
A= AdjA/det A
and Mj; is the determinant of the submatrix formed by deleting row 7 and
column j of A. In general the determinant of any submatrix is termed a
minor of A.
The following result due to Schur on evaluation of partitioned
determinants is often useful. If W, X, Y, and Z are matrices having
dimensions n x n,n X m,m xXnandm x m respectively, with W
nonsingular, then
det
K7|e
WX Wdet
(Z-YW!
X) (2.1
A similar result holds if Z is nonsingular (see Exercise 2.3).
Exercise 2.1.
What is the condition on the dimensions of A and B for the product AB
to exist? What does this imply for the matrices in eqn (2.5)?
Exercise 2.2.
If A and B are nonsingular matrices, use (2.5) to show that (A @BY ' =
A! @B’}. What is the corresponding expression for (AB) *?
Exercise 2.3.
By considering the product
w' o| |wx
-Yw' Fry oz
det u WX]
A =
_ det Z det (W—XZ
NeoSs Y).
Hencededucethat
det U, AY )= det UZ, —FA)
Exercise 2.4.
Show that the matrix equation
AX +XB=C
Preliminary matrix theory pag!
ay 12 Gin
a 422 Zon
ay = . ’ a> a > » Ayn = 4 (2513)
with the a;; allowed to be complex in general. The set of all such vectors
is called an m-space, denoted by C” (or R™ if the aj; are restricted to be
real). Both R™ and C”™are examples of a vector-space, and development
of the theory of vector spaces can be found in standard texts on linear
algebra. Notice that the vectors in (2.13) could equally well be written in
row form.
ipa, X;, are scalars then the vector
Example 2.1. If
1 6 =|
a, = 9) : ay = 0 > a3= ! — (2.15)
3 2 2
28 Preliminary matrix theory
1
a, eG, a4 + 243
or
a, =—2a,+ 4a;
Or
1 1
a3 81 +74:
Thus each vector can be expressed as a linear combination of the other
two, sOa,, 2, a3 are linearly dependent, whereas any two 4; , a; are
linearly independent.
Given a particular set of n column m-vectors in (2.13), then the set of
all linear combinations (2.14) forms a subspace of C™. This subspace is
said to be spanned by a,,...,4y. If every column m-vector can be
expressed as a linear combination of the a; then (2.13) is termed a
spanning set for C™.
1 0 G 1
aie 0 Sa 1 qe tiie 6) s @a— 1 (2.16)
0 0 1 1
x}
Xx ,X1,Xq arbitrary.
0
The vectors a,, 42,43, 44 forma spanning set for C*, and for example
In general there will not be a unique spanning set for C”. However, if
the members of a spanning set are linearly independent then the set is
called a basis, and it can be shown that any basis for C” contains exactly
m vectors, so that C™ is said to have dimension m.
Preliminary matrix theory 29
Example 2.3. For the vectors in (2.16)a4 =a, +a, +a3 but
0 0 0
As was seen in Example 2.1, any two of the columns are linearly inde-
pendent, so rank A = 2. Similarly, any two rows of A are independent.
The rank of A can be shown to be equal to the order of the largest
nonsingular submatrix of A. In particular, if A is square then it is non-
singular if and only if its rows (columns) are linearly independent.
Suppose now that the a;; are the coefficients in a set of m linear
algebraic equations in n unknowns
n
DS iy AB i 1,2, cs Mt (2.18)
jal
These equations can be written in matrix-vector form as
Ay =b (2.19)
30 Preliminary matrix theory
where
x=[x1,X0,0
0bXelt)b=bi, bose-oe (2.20)
Notice that the elements of Ax in (2.19) are precisely the same as those
on the right hand side of (2.14). Thus the problem of solving (2.19) for
the x; is equivalent to finding the coefficients in the expression of 5 as a
linear combination of the columns a; of A, if possible. It then follows that
(2.19) possesses a solution if and only if
Rank
A=Rank[4,5] (2.21)
where [4,b] is the m x (m + 1) matrix obtained by appending b to A as
an extra column. Two particular cases should be mentioned: when A is
square the equations (2.19) have a unique solution if and only if A is
nonsingular, and when the equations are homogeneous (i.e. all b; = 0) then
the necessary and sufficient condition for a nonzero solution to exist is
Rank A <n. Similar remarks apply to the set of equations.
yvA=b, y= [1,)25---.¥ml -
Exercise 2.5.
If a given m-vector v is expressed in terms of a given basis, ie.» =X, a
+ ++++Xp7Qm , show that the coefficients x; are unique.
Exercise 2.6.
where the c; are scalars, deduce that the vectors b, Ab, A*b,... ,Ab are
linearly independent.
2.3. Polynomials
.
Preliminary matrix theory 31
Exercise 2.7.
Deduce that if there exist polynomials x(A) and y(A) such that
a(r)x (A) + BAYA) = 1
than a(A) and b(A) are relatively prime.
The converse of this result also holds.
Exercise 2.8.
If the coefficients a; in (2.22) are all real numbers, show that if c is a
root of a (A) so is its complex conjugate ¢
32 Preliminary matrix theory
(2.29)
-k,=ta=>Ni i=1
where tr A = ay, +dy) + ++++ Gny the trace of A.
Preliminary matrix theory 33
The solution of (2.25) is therefore a sum of terms w; exp (jt) and this
will be developed formally in Section 3.1. The w; are sometimes called
right characteristic vectors of A, the left vectors which are rows being
defined by :
vjiA==Api.
If the A; are all different from one another it can be shown that the w;
are linearly independent so the matrix
W= [Wi,W,; Hons»Wn]
0 1 0 0
0 0 1 . . °
Ga = ° e e. e e e (2.32)
34 Preliminary matrix theory
AX+XB=C (2.37)
Preliminary matrix theory 35
where X and C are n x m. Using (2.7) and (2.9) this can be written in the
form (see Exercise 2.4)
Dx =c (2.38)
which is a set of nm equations in the familiar forms (2.18). The solution
of (2.38) is unique if and only if the mn x mn matrix
D=A®@Im +I, @BE
is nonsingular. To find the condition for this to hold, consider
(In +€A) ® Um +€BT)=Iy BI +eD +e2A OB!
which has characteristic roots (see Exercise 2.9 (c))
(1 + €dj) (1 + euj) = 1 + (A; +uy) + €? Aju,
Exercise 2.9.
Prove the following
(a) (A")= (A) (6) 2(4*) =; ()
(c) An +A)=1+2AKA)= @_~—2GA™) = 1/2;(A),provided
A isnonsingular.
Exercise 2.10.
A square matrix U is said to be unitary if UU* =I. Use the preceding
exercise to show that IA,;(U) |= 1.
The result still holds when U is real, in which case it is called
orthogonal.
Exercise 2.11.
By considering eqn (2.30) show that if m is a positive integer the
characteristic roots and vectors of A” are Aj” and w;. Hence deduce that if
BOA)
=BoN™
+ByA™
| 46+++Bm
then the characteristic roots and vectors of B(A) are B(\;) and w;.
36 Preliminary matrix theory
Exercise 2.12.
Using the result of the preceding exercise, show that the matrix B(C),
where C is defined in (2.32), is nonsingular if and only if B(A)and k{A)
defined in (2.28) are relatively prime.
This shows that det B(C) is a resultant for the polynomials B(A)and
K(A).
Exercise 2.13.
If in the preceding two exercises m <n, show that the first row of
B(C)is
T= [Bi B= Le oesBixBox
0.0.020]e
Show also that the subsequent rows arerC, rC?,..., ross
(Hint: if e; denotes the ith row of J, first show that e; =e; -; C,
i=2,3,...,n,and then consider e;8(C)).
Exercise 2.14.
Show that if k, #0 then the inverse of C in (2.32) is also a companion
form matrix with first row
[—kn—1/Kkn»
Kn —2)has “meee» ki Kn, 1 [ky]
and deduce that the characteristic polynomial of C—!
is (N"/kn) k (1/A).
Using an argument like that in Exercise 2.13, show that if B(A) is the
polynomial in Exercise 2.11 with m <n, then the rows of B(C') are
sC"*! sC—-"*2 v2, sC~*, 8, where s = [0, .. . 50,Bo ae
Exercise 2.15.
If the characteristic roots of C in (2.32) are Ay, Az, ... , Ay, assumed
distinct, show that corresponding characteristic vectors are
Me [Ay Me ee |! Lf=12,
The matrix M= [m,,m2,..., my] is called the Vandermonde matrix.
Prove by induction on vn,or otherwise, that detM= TI (Aj —Aj).
far
Exercise 2.16.
The linear discrete-time system corresponding to (2.25) takes the form
x(k +1)= Ax (k);k=0,1,2....
Preliminary matrix theory 37
Show that a solution of the form x(k) = wAk leads to the same character-
istic root and vector equation (2.26).
Exercise 2.17.
Show that the matrix equation
AXB-X=C,
where A isn xXn and Bis m x m has a unique solution if and only if there
are no characteristic roots dj, u; of A, B respectively such that dj uy;= 1,
Li eee i ls2, ss «5 I.
A particular version of this equation occurs in Chapter 5 involving
stability of discrete-time linear systems.
Exercise 2.18.
By a suitable choice of Wand Z in Exercise 2.3, show that for any
matrices X(m x m) and Y(m x n)
clearly
¥(s)=C(sl —A)" B us)
= G(s) u(s) (2.43)
where the r Xxm matrix
G(s)=C(sl— A) 1B (2.44)
is called the transfer function matrix by analogy with the scalar case
(1.13), since it relates the Laplace transform of the output vector to that
of the input vector. A block diagram representation of the form shown in
Fig. 1.8 can still be drawn, with input and output now vectors, and the
operator is the matrix G(s) which of course reduces to g(s) whenr = m= 1.
Using eqn (2.34), the expression (2.44) becomes
hoes ge a ee Ps +
Example 2.5. Consider the electrically-heated oven described in
Example 1.2, and suppose that the values of the constants in (1.4) are such
that the state equations are
y=f[l O}x.
Preliminary matrix theory 39
Shooe
Or o ten a afl
oul
2 +35
using
(sf!
—A)'
=Adj
(s/—
A)/k(s).
In practice it may well happen that the differential equations describ-
ing a linear system may not all be first order, as they are in (2.39).
Suppose that there are &state variables £; and m control variables u;. After
Laplace transformation the equation corresponding to (2.41) is
T(s) £ (s) = U(s) u (s), (2.47)
d*x
m2aapee (x;—X2)
kpx.
40 | Preliminary matrix theory
t= B) T(s)
=ie =Ky+k,mS? =i
+k, +k, |
U6)= | 1
Assume that the output is y =xz, so that V(s) = [0, 1]. Hence from
(2.49)
where L(A) and M(A) are both polynomial matrices having determinants
which are nonzero constants (i.e. independent of A). It can be shown that
Preliminary matrix theory 41
where the elements in (2.51) are polynomials such that iz (A) divides
ix + 1(A), k= 1,2,...,7—1. It can also be shown that the iz (A) satisfy
where dx (A) is the g.c.d. of all minors of P(A) of order k (these minors
are of course polynomials) and dx is called the kth determinantal divisor
of P(A) (taking dp = 1). Notice that it easily follows from (2.52) that if
r=n then
tity ce ly =dy
= det P(A). (233)
Since the application of elementary operations to P(A) does not alter its
Smith form, the iz (A) in (2.52) are called the invariant factors of P(A).
As in (2.24), each iz (A) can be written
ip (A)=(A—0) Bk (A—
022)Bk,-- (A—04)PRS,
K=1,...,7 (2.54)
where the a; are in general complex numbers, and the notation is chosen
so that a, @2,..., Qs are the distinct roots of i, (A), which implies that
Br1, Br2, - - - Bryare all positive integers. In view of the divisibility prop-
erty of the iz, it therefore follows that all ae remaining 6xQwill be
nonnegative integers. The factors (A —aj) Bre occurring in (2.54) for all
values of j, k, 2 such that By.g#0 are tia the elementary divisors of
P(A). It can be shown that P; (A) and P, (A) are equivalent if and only if
they have the same rank and their respective sets of elementary divisors
are identical.
Example 2.7. If a matrix has Smith form
” 0 )
S=1]0 M4(A+ 1)? 0
0 0 AM(A+1)* (A+ 2)
NOt 1) OS Oa
Our main interest in this book is in the special case when P(A) is the
characteristic matrix MJ, —A. In this case the invariant factors are often
called the similarity invariants of A, because it can be shown that two
n X n matrices are similar if and only if their similarity invariants are
identical. This result enables us to determine when a given matrix A with
characteristic polynomial k(A) is similar to the associated companion
form matrix C defined in (2.32). The argument requires some further
definitions. Firstly, the Cayley—Hamilton theorem states in (2.36) that
k(A) = 0, but it can happen that there are other polynomials for which
this holds. Suppose that the monic polynomial m(A) is such that
If there are no polynomials which have degree less than p and satisfy
(2.55), then m(A) is called the minimum polynomial of A.
Example 2.8. If
it is trivial to verify that k(A) = \° and that A? =0, but A +m,/#0 for
any value of m,, so the minimum polynomial is m(A) = A’.
When m(A) is identical to k(A) then A is termed nonderogatory, other-
wise derogatory (much of the nomenclature in matrix theory was coined
in the Victorian era —the implication is clear that a derogatory matrix has
undesirable properties!). The next step in determining similarity of Aand
C is to note that it can be shown that m(A) is equal to the similarity invari-
Preliminary matrix theory 43
oS) onl
b= kOenent.
¢ aOels (2.57)
‘Shine eal
the determinantal divisors are d; = 1,d, =(A— 1), d3 =(A—1)?. Thus
eqn (2.52) gives
BepiloigaN) pis=(A-1)’,
and m(A) =i3. Clearly k(A) =(A—1)° so A is derogatory, and there is no
nonsingular matrix 7 such that
Exercise 2.19.
Derive the transfer function in Example 2.6 using the first order linear
state equations obtained in Exercise 1.2.
Exercise 2.20.
Suppose a harmonic input u = ugcos wt, where ug is a constant vector,
is applied to (2.39). Assume that x(t) has the form acos wt + bsin wt,
and determine the constant vectors a and b (assume A has no purely
imaginary characteristic roots). Hence show that the output (2.40) is
y= Re{G(iw)}uocoswt —Im{G(iw)}uo
sinwt
where G(s) is defined in (2.44).
This provides an interpretation for the frequency transfer function
matrix G(iw), and when G is a scalar the expression for y reduces to
precisely that given in eqn (1.20).
Exercise 2.21.
Show that the minimum polynomial of any square matrix A is a factor
of its characteristic polynomial. (Hint: assume that division of k(A) by
m(A) produces a remainder term, and obtain a contradiction).
Exercise 2.22.
Given ann x n matrix A, if a column n-vector b can be found such that
diag Ym, (Ar), Im, Ar), - «+Amy Ar) Sn, Ar), +++Ft, Ag) I], (2.59)
using the notation of (2.3), where Jz (A) is the k x k Jordan block
r 1 0 . 0 0
0 r 1 : 0 0)
Bein) *| : : 2 : é (2.60)
0 0 : r 1
0 0 , ° 0 r
(an alternative form has the 1s below the principal diagonal). The distinct
characteristic roots of A are \y,2,..., Aq and the multiplicity of A; is
m, +m ++ + +mg, and so on. The Jordan form of a matrix is unique up
to the ordering of the blocks in (2.59), and can be written down once the
' elementary divisors of J —A are known (see Exercise 2.24). If none of
the characteristic roots of A is repeated, so that g =n, eqn (2.59) reduces
to the diagonal matrix (2.31). In this case each Jordan block is simply
J; (Ai) =);
It should also be noted that the Jordan form of A can be a diagonal
matrix even if some of the characteristic roots of A are repeated. There
is just one linearly independent characteristic vector of A associated with
each Jordan block, so the total number of such vectors is equal to the
number of blocks in the Jordan form (2.59). In other words, A is similar
to a diagonal matrix if and only if it has n linearly independent character-
istic vectors. An important special case of this is that every real symmetric
matrix (i.e. Al = A) has a diagonal Jordan form; in addition a transforming
matrix can be found which is orthogonal.
Exercise 2.23.
It will be shown later (Section 2.8) that if A is a real symmetric matrix
then all its characteristic roots are real. If none of these roots is negative
deduce that there exists a nonsingular matrix T such that
Tee
Exercise 2.24.
Show that \/x, —Jx(Aj) has the single elementary divisor (A —dyihe
It can be shown that the elementary divisors of \J —A are
Exercise 2.25.
Prove by induction on r that
we(AK elie
1 2
J, QO)?
=| 0 ne Arye 6 9230:4:
Nao
0 0 nN
WAlle=|
> >, lay? (2.62)
where |a;| denotes modulusof aj;. For a vectorx (2.62) gives
1
n
Ixle=|_
> bel?
rT
(2.63)
48 Preliminary matrix theory
which can be thought of as the ‘length’ of the vector x. Other norms can
and if 2 is a vector
Rim || A, -A || =0
T — co
for any norm. It is easy to show that if Pand Q are constant matrices
then PA,Q > PAQasr > ©, In particular a square matrix A can be
written as A= 7JT~' where J is the Jordan form in (2.59). Thus A” =
TJ’ T~* and since the elements of J’ depend upon powers of the charac-
teristic roots A; of A (see Exercise 2.25) it follows that Rim A” = 0 if and
r = yee
only if all the characteristic roots of Ahave modulus less than unity.
co
The infinite matrix series > Ax converges provided the sequence of
k=0
r
partial sums X; = Azx,r=0,1,2,..., converges to a finite limit as
k=0
co
r > ©, The series is absolutely convergent if the scalar series > || Ax |
k=0
is convergent, and absolute convergence implies convergence.
Preliminary matrix theory 49
A powerful general result states that the matrix power series in (2.70) is
convergent if all the characteristic roots \j of A satisfy | Aj |<R.
converges provided A is such that all its characteristic roots have modulus
_ less than unity.
It can be shown that when all the ); are distinct, then assuming con-
vergence,
k=1
where
the same matrix function. Note also that (2.71) still holds if f(A) is a
finite series.
An alternative way of evaluating f(A) when all the Aj are distinct is as
follows. As in (2.23) we can write
which are 7 linear equations for the n unknown coefficients 7;. Examples
using both approaches to evaluation of f(A) are given in Chapter 3.
When A has repeated characteristic roots the general expression for
f(A) is
f(A)=» [FOue)
Zier+£de) Zia+2 +fK—YD
Ag)Zkag]
k=1 (2.76)
where the a; are the indices in the expression for the minimum polynomial
in (2.61). In (2.76) f (Ax) denotes the rth derivative of f with respect
to A, (not t) evaluated at \ = Ax. The matrices Z,; are constant and again
are determined entirely by A, not the function f. Although an explicit
formula can be given for these matrices, they are best determined by
taking simple choices for f, and again this is illustrated by means of
examples in Chapter 3. When all the A; are distinct so that q =n, (2.76)
reduces to (2.71). If some of the ); are repeated, but the minimum poly-
Preliminary matrix theory 51
nomial m(A) has only simple roots (i.e. in eqn (2.61) all a; = 1) eqn
(2.76) becomes
q
f(A)= >, Zi fx) (2.71)
k=1
where the Zz are defined as in (2.72) except that the upper index in the
product is q instead of n.
We end this section with a warning: although the power series (2.69)
has been used to define a function of a matrix, it cannot be invariably
assumed that properties of the scalar function carry over to the matrix
case (see for example Exercise 2.27).
Exercise 2.26.
For what matrices A does the series for sin A converge? Use (2.71) to
determine sin A if
Exercise 2.27.
If A and B are twon x n matrices prove that
Exercise 2.28.
Consider an (n —1)th degree polynomial p(A) which takes the values
PAK) = px, k =1,2,...,n, with all the Ay distinct. By writing
Exercise 2.29.
Show that in eqn (2.71) the matrices Zz satisfy
n
ee n
>) Ze 47,7 1,2,352.
k=1 k=1
2.8. Quadratic and Hermitian forms
non
h(x) =x*Ax = > ‘ ayxix;, (2.79)
i=1 j=l
with A now a Hermitian matrix (A* = A). Notice that the complex con-
jugate of h is
h=x!Ax=(x*A*x)?
=h,
showing that h is real for all vectors x. We can show similarly that all the
characteristic roots Aj of A in (2.79) are real. For let w; be a characteristic
vector satisfying (2.30), so that
real, and the argument still applies if Ais symmetric. Because of this
property algorithms for calculating characteristic roots of symmetric and
Hermitian matrices are simpler than for general matrices. Also, as we
remarked in Sections 2.6, every real symmetric matrix is orthogonally
similar to the diagonal matrix of its characteristic roots, and this still
holds for Hermitian matrices except that the transformation matrix is
then unitary.
In our work on Liapunov stability theory in Chapter 5 we will be
interested in Hermitian or quadratic forms v(x) which do not change sign
as x varies, so the following definitions are needed:
(i) v(x) is positive (negative) definite if v(0) = 0 and v(x) > 0 (K0)
when x # 0.
(ii) v(x) is positive (negative) semidefinite if v(O) = 0 and v(x) 2 0
(< 0) for all values of x, with at least one x #0 such that
v(x) = 0.
n
> wl P. (2.83)
i=1
Since U is nonsingular (see Exercise 2.10) y = 0 if and only if x = 0, so
it follows that v will be positive (negative) definite if and only if all
54 Preliminary matrix theory
Xi (A) > 0 (< 0); positive (negative) semidefinite if and only if all \; (A)
20 (<0), with at least one A; (A) = 0; and indefinite if and only if at
least one A; (A) > 0 and one ); (A) < 0. In fact it is possible to avoid
calculation of the Aj by reducing v(x) to a sum of squares by Lagrange’s
method (Mirsky 1963, p. 371), the numbers of positive, negative and zero
coefficients in the sum being the same as in (2.83).
An alternative approach involves the principal minors P; of A, these
being any 7th order minors whose principal diagonal is part of the principal
diagonal of A. In particular the Jeading principal minors of A are
1 ps 3
A= |2 4 6
3 6 9
and the principalminorsare
ge: a cated
A2 4ae =0, | 4oa 6 =0, ia9 pa Lents
Exercise 2.30.
A skew symmetric real matrix S is defined by ST =-§.Ifq=x1Sx show
by considering (q)! that q =0 for all vectors x. Similarly show that x*Sx
=0 if S is skew Hermitian, i.e. S* =—S.
Exercise 2.31.
Show that any real matrix A can be written as A= A, + A, where A, is
symmetric and A, is skew symmetric. Hence using the result of the pre-
ceding exercise, deduce that x! Ax =x! A,x, allx.
A similar argument applies for complex matrices.
Exercise 2.32.
Let A be a Hermitian matrix, and write it as A =A, + i42, where A,
and A, are real and i=+/(— 1). Show that A, is symmetric and A is
skew symmetric. Hence deduce that if A is positive definite, so is A,.
Exercise 2.33.
Two column n-vectors a and b are said to be orthogonal if a*b = 0. Show
that if w;, w; are characteristic vectors associated with distinct charac-
teristic roots of a Hermitian matrix A then w; and wy;are orthogonal.
Exercise 2.34.
If A is an x n symmetric matrix having all aj; = 1, deduce that it has a
single nonzero characteristic root equal in value to n (Hint: use (2.29)).
3 Matrix solution of linear systems
’
We were able to express the linear examples in Section 1.1 in the matrix
form (1.1). To begin with we shall consider systems without the presence
of controlled variables, and in this and the following two sections we
discuss methods of finding the solution for the state vector x (¢) of the
nth order system described by
x = Ax, (3.1)
x(0)=Xpo. (3:2)
We shall first assume that all the characteristic roots Ay, A2,.--, An
of A are distinct. In fact in real-life situations this is not too severe a
restriction, since if Adoes have repeated roots, very small perturbations
in a few of its elements (which will only be known to a certain degree of
accuracy) will suffice to separate these equal roots. If w; is a character-
istic vector corresponding to Aj then W,, W2,...,W» are linearly
Matrix solution of linear systems 57
Lc; =A LCjW;
= DAiciw; ,
n
x(t)= 3 ¢; (0) exp Cyt); - (3.4)
ped
This generalizes our informal remarks in Section 1.2 on the solution of
(1.7). If W denotes the matrix whose columns are wy, W2,..-, Wy then
it is a standard result that the rows v,,2,...,, of W™ are left
characteristic vectors of A. Since we have v;w; = 1, v;w; = 0, i #/, multi-
plying (3.4) on the left by v; and setting ¢ = 0 in the resulting expression
gives v;x (0) = c;(0). Thus the solution of (3.1) satisfying (3.2) is
x()=>, @1x(0))
exp(it); (3.5)
i=]
Notice that the expression (3.5) depends only upon the initial conditions
and the characteristic roots and vectors of A, and for this reason is referred
to as the spectral form solution (the set {rj }being the spectrum of A).
58 Matrix solution of linear systems
1 1
xXy(t) 1 1
= (2x1 (0) +x (0))e* -(x, (0) +x
X(t) =I =)
(3.6)
Exercise 3.1.
Find the general solution of (3.1) subject to (3.2) in each of the follow-
ing cases:
(a) -] -] (b) 1 0 -]
A=
2 —4 Ae at 2 1
2 2 3
Exercise 3.2
If n = 2 and
nO= ve" when x(O)=| 1
It ee “2
x(t)= | e?! when x(0)=]| 1
Lieet 2,
Matrix solution of linear systems 59
find the general solution of (3.1) subject to (3.2) using the linearity
property described in Chapter 1. Hence find also the matrix A.
x(¢)=exp(Af)xo (3.7)
Wedefinetheexponential
matrixby
exp(Af)=1+tA +(t?7/2!)A?+(2/3) Ar tees. (3.8)
The topic of infinite matrix series has already been discussed in Section
2.7 and it follows from (2.69) and (2.70) that since exp (zt) converges
for all finite scalars z and ¢ then the series on the right in (3.8) converges
for all finite ¢ and all x n matrices A having finite elements. It is clear
from (3.8) that exp (0) =/ and that
@(t,f9)=exp[A(t—
t)] (3.11)
is called the state transition matrix, since it relates the state at any time rf
to the state at any other time fo. It is left as an easy exercise for the reader
60 Matrix solution of linear systems
to verify using (3.11), (3.8) and the result of Exercise 3.4 that ®(¢,to) has
the following properties:
d
Pl fo)=APCF,
fo)
&(t, A =1 (3.12)
Evaluation of exp (Ar) when all the A; are distinct can be achieved by
Sylvester’s formula (2.71) which gives in this case
n
exp (At)= > Zq exp (Ags) (3.13)
k=1
where
n
Zp = Wy (A —D/O —y)- (3.14)
]=
J#K
Example 3.2. Using the matrix in Example 3.1, eqn (3.14) gives
2 1
Z, =(4-(-2))/C1 - €-2)) =
>) sd
oI =
pata (1) )/iC2-—C1)) =
2 2
so that rp =e" —e
7", r, = 2e* —e?". Hence from (3.15)
x(t)= ex,
SXotXo=Ax;
orafterrearrangement
x(s)= (sl —
A)" Xo,
sothat
x(th=£' {(sI-A)! xo}. (3.17)
Bycomparison
with(3.7)it thenfollowsthat
£-{ (sl A)"}= exp(Ad), (3.18)
which is a generalization of the well-known result when n = 1.
62 Matrix solution of linear systems
Exercise 3.3.
Prove directly from (3.8) that exp(4 +B) # exp A exp B for two
square matrices A and B unless Aand Bcommute with each other.
Exercise 3.4.
Using the result of the previous exercise, determine the inverse of exp A
and hence deduce that exp A is nonsingular for any square matrix A.
Exercise 3.5.
If A= diag[a,,a2,...,a,] show that
expA 7 diag [exp(@,), exp(a), ste, exp(an)] :
Exercise 3.6.
Consider the equation of simple harmonic motion 2 + w*z = 0. Take as
state variables x; =z and x2 = 2Z/w,and find the transition matrix @(t,0)
using (3.11) and (3.8). (Hint: Find A”).
Exercise 3.7.
Find the general solution of (3.1) and (3.2) using (3.13) or (3.15) for the
two cases where A is given in Exercise 3.1.
Exercise 3.8.
Find the general solution of (3.1) subject to (3.2) when A is the matrix in
Exercise 3.1 (@)by determining (sJ -—A)’ and using eqn (3.17).
Exercise 3.9.
Use the exponential matrix to solve the rabbit-fox environment problem
of Exercise 1.4 subject to the same condition as in Exercise 1.7, namely,
Qa;/a3 =a2/A4.
Exercise 3.10.
Consider the expression for exp(Af) in eqn (3.13). By taking Laplace
transforms of both sides, or otherwise, show that
n
ZK= Ln
k=1
AZ = AZ, ZEA=AgZE,K=1,2,...,0
Matrix solution of linear systems 63
and
Lif; = §5/2;
Exercise 3.11.
Write Jz (A) in (2.60) in the form XN+ K, and hence show that
eee Pe wee ik 1)!
] t .
exp{Jx(A)}t=exp(Ar)
op.)
Exercise 3.12.
‘Verify that the solution of the matrix differential equation
a AW(t)+
W(1)B,
W(0)=C,
where A and B are square and constant, is W= exp(4t)C exp(Bt).
Since
A 4 WB)
W-A= |. -1 r 11
OF AO Na
it is easy to verify directly from the definition (see Section 2.5) that the
determinantal divisors of WW’ —A are d; = 1,d, =A—2,d3 =(A—-2)? so
by (2.52) the similarity invariants of A are i; = 1, i, =d,/d, =A-2,
i; =d3/d, =(A— 2)’, this last polynomial being the minimum poly-
nomial m(A) of A as stated in Section 2.5. Hence g = 1, Ay = 2, a, =2
in (2.76) which becomes
and this expression holds for any function f(A). If we choose for simplic-
ity f(A) = 1 and f(A) =A —2 then this gives in turn Z,, = and Z,) =
A —21. Finally, taking f(A) = exp(Af) in (3.19) produces
afer? +(A—21)te?!
LP ye —4t —22t
0 0 1
Exercise 3.13
Find the general solution of the system
¥1=—2x, +x,
Xa Saad —4x,
Exercise 3.14.
Evaluate exp(Ar) when
g pa ]
am Rae 0 ]
1 “A 2
X¥=Ax + Bu (3.20)
x(t)=exp
(Ar)
[xo+ffexp
(—Ar)
Bu(nr| ; (3.22)
If the initial condition is x (to) =xo, integration of (3.21) from fo tot
and use of the definition of @in (3.11) gives
x(t)
=®(t,
f)xo+i (to,7)Buca)ar
|. (3.23)
Thus if u(t) is known for t > to, x(t) can be determined by finding the
state transition matrix and carrying out the integration in (3.23).
z(t) being the displacement from some fixed point. In state-space form,
taking x, =z and x2 =Z as state variables, this becomes
d x4 (t) 0 1 xX4(t) 0)
at 5 s u(t)
xX2 (t) 0) 0 X2 (t) 1
= Ax + Bu, say.
SinceherewehaveA*=0,(3.8)reduces
toexp(At)
=/ +At,so(3.22)
gives
x, (¢) Fre x (0) i Jost ip | 6 Ora) ar.
x2(0) 0 1| |x. (0) 0:1). 4°. |Oudanaae
Solving for x, (t) leads to
Exercise 3.16.
Determine exp(Af) for the electrically heated oven problem of Example
1.2, assuming that the state equations take the numerical form given in
eqn (2.46) in Example 2.5. Suppose that initially the oven interior and
jacket are at room temperature 7, and that at t = 0 the heating element
is switched on, providing a constant input of magnitude 2 units. Use eqn
(3.22) to find an expression for the subsequent inside temperature of the
oven.
Exercise 3.17.
Consider the system (3.20) with
Matrix solution of linear systems 67
and take u(t) to be the unit step function defined in (1.16). Evaluate
exp(A?) using the power series definition (3.8) and hence show that the
solution of (3.20) subject to x(0) = [1, 0]? is
1+3t+—t oy + eee
dL
mattuloss
ere et
Exercise 3.18.
Consider X= (A + €a)x where A and ware constant n x n matrices and
€ is a parameter. Using (3.22) show that if € is small, then a first approx-
imation for x(t) is X(t)xo where
X(t)
=exp
(Af)
E+€i exp
(—Ar)a
exp
(And
Show also that the next approximation produces an additional term of
order €?.
THEOREM 3.1. If A(¢)is continuous for ¢ > 0 then (3.25) has a unique
solution for t 2 O given by x(t) = X(t)xo9, where X(f) is the unique n x n
matrix satisfying
dyY_ hy
Wt YA(t), Y(O)=/.
< (YX)
=vr+yXx
=— YAX + YAX
=0
is the solution of (3.25) with initial condition x (to ) = x9. The expression
(3.28) has the same form as that for the time invariant case, given in (3.10),
although of course the transition matrices involved are quite different.
However it is most interesting that although in general it is not possible
to obtain an analytic expression for the solution of (3.26), and therefore
for ®(t, fo) in (3.27), this latter matrix possesses precisely the same
properties as those for the constant case given in equations (3.12). This
Matrix solution of linear systems 69
x(t)=P(t,
fo)K+i P(to,
T)B(7)
u(r)
ar| (3.30)
where® is definedin (3.27).
Proof. Using the standard method of variation of parameters, put
x = X(t)w(t) where X(f) is defined by (3.26). Substitution into (3.29)
produces
axe
qp AAwtx
dw
=AXw + Bu.
Hence
dw _
X(t)Gp=Bu
so
dw =i
EIR — t B
Tees (t) Bu
which on integration gives
The desired expression then follows using x9 = X(to )w(to ) and (3.27).
The development in this section shows that some of the results on
linear systems carry over even when the matrix elements are time varying.
This is clearly a useful aspect of the state space approach, since transform
methods can only be applied to equations with constant coefficients.
70 Matrix solution of linear systems
Exercise 3.19.
Show that when n = 2 in (3.26),
Exercise 3.20.
Verify using (3.27) that the properties of the transition matrix listed in
(3.12) do indeed carry over to the time varying case.
Exercise 3.21.
If B(t) = f’ A (r)dr, verify that the solution of (3.25) is x(¢) = exp (B()) xo
provided B(t) and A(t) commute with each other for all > 0.
Exercise 3.22.
Show that if ® is defined by (3.27) then [®~' (t, fo)]" is the transition
matrix for the system Z = —A!(¢)z (the adjoint system).
Exercise 3.23.
A system is self-adjoint if A' (t) =—A(t). Show that (¢, fo) in (3.27) is
then an orthogonal matrix.
Exercise 3.24.
Verify that the solution of the matrix differential equation
ae A(t)W(t)h+W(HAT(D),
W(to)
=C,
is W(t) = P(t, to) CB! (t, to), using the properties of ® which is defined in
(3.27).
Compare with Exercise 3.12.
where x(k), u(k) denote the values of the state and control vectors x(KT)
and u(kT) respectively (k =0,1,2,...). The similarity with the contin-
uous-time equation (3.20) is obvious, and the dimensions of A and B are
the same. We now develop matrix methods for solution of (3.31). Consider
first the situation when there is no control and A is a constant matrix, so
(3.31) becomes
Clearly
x (Koae 1) = Ax (ko)
x (Ko + 2)=Ax(Ko + 1) =A?x (Ko), etc.,
Ak =roA try
and since the characteristic roots of Aare —1 and —2, eqn (2.75) gives
Ameer HE 2) 42,
where the Z; are identical to those found in Example 3.2. This emphasises
the fact previously mentioned that the Zz in Sylvester’s formula (2.71)
are dependent only onA, not the particular function f(A) being evaluated.
WhenA has repeated roots the formula (2.76) can again be used.
72 Matrix solution of linear systems
for any function f(A). Setting f(A) in (3.34) equal to 1 gives Z,, =J, and
f(A) =X +2 gives Z,. =A + 2/. Then taking f(A) = \™ gives
O(k,k) =I
(3.37)
&(ko,k) =®—! (k, ko), provided A is nonsingular
Now return to (3.31) but assume A and B are still time invariant. We
have
k-1
=A4k-kox,
+> Ak-i-1Bu(i), (3.38)
i=k,
=&(k, k-1
ko)E +y ®(Ko,
i+1)Bu| (3.39)
i=k
,
_using (3.35) and (3.37). The expression for x(k) in (3.39) corresponds to
that in (3.23) for the continuous-time case.
In the case when A and B have time varying elements the development
is a little easier than for the differential equation system in Section 3.5.
We can write
X(ko + 1)=A(Ko T)X0 ,
X(ko + 2)=A[(ko + 1)T] x(Ko + 1)
=A [(ko + 1)T]A (KoT)Xo0,
k=1
O(k, ky)= I] A(iT). (3.40)
i=k
I=Ko
i=k,
v4 Matrix solution of linear systems
Exercise 3.25.
Find A’, when A is the matrix in Exercise 3.1(@),by using the expres-
sion (2.71).
Exercise 3.26.
By choosing as state variables x, (k) = X(k), x2 (k) = X(k + 1), write the
scalar difference equation of Example 1.7 in the form (3.32), and find
®(k, 0).
Exercise 3.27.
Find ®(k, 0) for the system having the A matrix given in Example 3.6.
Exercise 3.28.
The Fibonacci numbers are 0, 1, 1, 2, 3,5, 8, 13, ... each number in the
sequence being the sum of the preceding two. These integers arise naturally
in a surprisingly large number of biological and other situations (see for
example Holland (1972) for an elementary account). Let X (k) denote the
(k + 1)th number, and write the difference equation for X in the form
(3.32), as in Exercise 3.26. Find ® (k, 0) and hence show that
X(k)
=E+(5)s
+(5)5°
+(7)5°
ee| jok-1,
where
(; )denotes
theusual
binomial
coefficient.
Exercise 3.29.
By using z-transforms to solve (3.32), deduce using (3.33) that
Z'{z@i-A),1}=A*.
This is the result corresponding to (3.18), and is the matrix generalization
of Example 1.6(b).
Matrix solution of linear systems WS
We have assumed so far in this chapter that the linear system equations
are in vector-matrix form. In fact as we have seen in Chapter 1, classical
linear control theory deals with scalar differential or difference equations
of the forms (1.7) and (1.26). Consider a simplified form of (1.7):
w=Cw+du
where
1 0 0
0 1
cC=| - : : ; , : ; ES
1
ohn —kn—-1—kn-2 * ; eras.
Wa5(Wytis fyo. wy
and
Asal ONO
creer:OS |g (3.46)
x = Ax + bu (3.48)
be put into the classical form (3.42), or its matrix equivalent defined by
(3.43) and (3.44)?
Example 3.7. We saw that in Example 1.2. concerning the electrically
heated oven the equations describing the system arose naturally in the
matrix form (3.48). Suppose that the values of the constants in (1.4) are
as in eqn (2.46) so that eqns (1.2) and (1.3) become
ope apSpeeaie
Hence
Koitise meee
which has the desired form. We would like to determine when and how
such a procedure can be carried out in general. The answer to our ques-
tions is provided by the following result:
[!tA
tA?
TS | (3.51)
Lane)
where ¢ is any row n-vector such that 7 is nonsingular, assuming for the
present that at least one suitable ¢ exists. Denote the columns of J’ by
S1,S82,...,5, and consider
tS tS $ * Sy, ]
tAs, tAs, : ° tds, ] @)
for some scalars a;. Multiplying (3.54) on the right by b and using (3.52)
givesa, = 0. Similarly, multiplying (3.54) on the right successively by
Ab, A?b,..:,A"~"b gives a, —1 =0,@,—2 =0,..., 0, =O, establish-
ing independence.
To prove the converse of the theorem, if a nonsingular T exists then
rank [b, Ab,...,A"—!b] Stank [Tb, TAB...» eee)
and it is easy to verify using (3.45) and (3.46) that this last matrix has
the triangular form
ie Ove, 0) ne 6°74
L 6) 1
6, 0,
2 s< by
; On —?
On-1
il
0:=-> K+) Opeju{OP eo oe ee =,
J=0
=
Matrix solution of linear systems 79
kn-—te thy 2. f ky ]
ee) 1, eae ° 1 0
(Ove = c : 3 6 : . . (3.58)
ky 1 je) S)
1 0 (ee) (3
Finally the inverse of 7 is given explicitly by
Toray (3.59)
Example 3.8. Find the matrix T and the transformed system when
1 =3 1
A= b=
4 2 1
=] 1
jp ane
3 5
and
es) 1
:Ric
3 1
A simple calculation gives
0 1
TAP =
—14 WwW
80 Matrix solution of linear systems
or
2-324 147 =H;
x=Ax (3.60)
Y=Cx, (3.61)
where A is ann xXn matrix as before, c is a constant row -vector and
y(t) is the output variable.
where
0 0) en
l 0 ~€n—]
0 1 F c :
E= (3.63)
0) 0) : 1 —e
and
f= (0; 0,04 231), (3.64)
Matrix solution of linear systems 81
provided
cA?
rank =n. (3.65)
seat
Conversely, if a nonsingular P exists then (3.65) holds.
Proof. This follows along similar lines to the proof of Theorem 3.4, so
we do not give all the details. The transformation x = Py when applied to
(3.60) and (3.61) produces »= P"'APy, y = cPy so we first require to show
that P-'AP has the form (3.63). This is achieved by setting
RalroArs .. ae lp (3.66)
assuming that some column n-vector r exists which makes P nonsingular.
If the rows of P"' are denoted by q,,q2,...,4n, then comparison with
_the elements in the identity P-' P=J establishes that P~'! AP has the
desired form (3.63) with last column e =— gy —; + ;A"r,i=1,2,...,0n.
The second requirement, that cP = f, holds because of (3.65), and nonsingu-
larity of Pis shown by proving that its columns are linearly independent.
The proof of the converse is also similar to that in Theorem 3.4. The
matrix P is constructed using (3.66) and the condition cP =f.
Notice that F is also called a companion form matrix because its
characteristic polynomial is
X(k+n)t+kX(k+n-1)+++++ky,X(k)=u(k) (3.68)
the only modification required is that (3.43) is replaced by
wy,(k) =X(k), w2(k) =X(K+1),° °° , Walk) =X(kK+n-1) (3.69)
82 Matrix solution of linear systems
so w,(k + 1) = Ww;
+ 1(K), and (3.68) is equivalent to
w(k + 1) = Cw(k) + du (k) (3.70)
where C and d are precisely as defined in (3.45) and (3.46). We have in
fact already used a simple case of (3.69) in Exercises 3.26 and 3.28.
Theorems 3.4 and 3.5 now carry over with no changes other than that
(3.44) is replaced by (3.70), and similarly in (3.48), (3.60), and (3.62)
x(k + 1) and v(k + 1) replace derivatives.
Exercise 3.30.
Find the matrix T in Theorem 3.4 and the transformed system for each
of the cases where the matrices are as given in Exercise 3.1 and the cor-
responding vectors b are
an velo
Exercise 3.31.
Two platforms P; and P, are connected to each other and to a fixed
support by springs and dampers as shown, and a control force u(f) is
applied to Py.
|
Spring constant =|
Damping
coefficient =k,
P,
Spring constant = 1
Damping
coefficient =k,
FIG. 3.1
Matrix solution of linear systems 83
The forces exerted by the dampers are proportional to velocity, and the
springs obey Hooke’s law. Assume that the masses of the platforms and
the other components can be neglected, and that the platforms remain
horizontal. Derive the state equations in the form (3.48), taking the
displacements x; and x, from equilibrium as state variables. Find the
condition to be satisfied by k, and k, for there to exist a nonsingular
transformation of variables which puts the equations into the form
(3.44). If k, = + k, = * determine T and the transformed system.
Exercise 3.32.
Find the matrix P in Theorem 3.5 and the transformed system when
1 p 0
A={3 rail es c= [0,0, 2].
0 4 0
Exercise 3.33.
Prove the converse of Theorem 3.5.
| Exercise 3.34.
Verify that when A is a scalar multiple of the unit matrix then it is
not possible to transform it by a nonsingular transformation into either
(3.45) or (3.63).
Exercise 3.35.
Show that if
Exercise 3.36.
If V denotes the matrix in (3.65), show that the inverse of the trans-
formation matrix in Theorem 3.5 can be written as P~' =(U)"'V,
where (U) ! is as stated in (3.58) but with the k; replaced by the e;
in(3.67),i=1,2,...,n-1.
4 Linear control systems
4.1 Controllability
implies that the definition holds for all x9 and x¢, and several other
types of controllability can be defined—for example, complete output
controllability requires attainment of arbitrary final output. The control
u(t) is assumed piecewise continuous in the interval fo to f,, that is,
continuous except at a finite number of points in the interval.
JX, =u
X2 = a3X>2.
xqeid(ty,
te)[xo
+(x©(to,
7)
B(r)u(r)
ar| (4.3)
86 Linear control systems
Oxianeh)
fx.—H(to,
er}+f' &Co,
DBOuC)
ar|
Since ® (f,, fo) is nonsingular it follows that if u(t) transfers x9 to x¢,
it also transfers xy —®(fo, t,) Xpto the origin in the same time interval.
Since x and x¢ are arbitrary, it therefore follows that in the definition
the given final state can be taken to be the null vector without loss of
generality.
For time invariant systems the initial time to in the controllability
definition can be set equal to zero, and a general algebraic criterion can
be derived:
THEOREM 4.1. The constant system
x = Ax + Bu (4.4)
(or the pair [A, B]) is c.c. if and only if the n x nm controllability
matrix
U= (BAB A2B,..5A ae (4.5)
has rank n.
Proof.
Necessity. We suppose (4.4) is c.c. and wish to prove rank U =n.This
is done by assuming rank U <n, which leads to a contradiction. For then
there will exist a constant row n-vector g # 0 such that
qB=0,qAB=0,...,qgA"
1B=0. (4.6)
In the expression (3.22) for the solution of (4.4) subject to x (0) = xq set
t=f,, x(t,) = 0 to obtain, since exp(Afr,) is nonsingular,
xg=f° exp-An)Bu()
ar. (4.7
Now as in (3.15), exp(—Ar) can be expressed as some polynomial r(A )
in A having degree at most n —1, so (4.7) becomes
exp(—Aty
)x(t;)=X +(soB+s,AB+++++s, ,A"-'B). (4.10)
Since rank U=n it follows that for any given xq it will be possible to
choose the s; (and hence by implication u(7)) so that the right hand side
of (4.10) is identically zero, giving x (¢,) = 0. This establishes that (4.4) is
c.c. since the conditions of the definition are satisfied.
COROLLARY 4.1. If rank B = p, the condition in Theorem 4.1 reduces
to
rank [B, AB,...,A"~PB] =n.
Proof.Definethematrix
Peete AR. LA BYR =0,12yacin-
‘If rank Up = rank Up, it follows that all the columns of A&*?B must
be linearly dependent on those of Up. This then implies that all the
columns of A£*?B, A**3B, . . . must also be linearly dependent on
those of Up, so that rank Ug = rank Up,, = rank Ug, =.... Hence
the rank of Ux increases by at least one when k is increased by one, until
the maximum value of rank U; is attained when k = 2. Since rank Up =
rank B = p, and rank U; <n it follows that p + &£ <n, giving <n -—pas
required.
ae! 2 1
= x+ Uu. (4.11)
1 eal 0
The controllability matrix (4.5) is
1 =2
=
0 1
88 Linear control systems
which has rank 2, so the system is c.c. That is, it is possible to control
(in the sense of our definition) the temperatures of both the oven interior
and the jacket by means of the heating coil in the jacket. However, this
does not tell us how the control can be applied to satisfy given objectives.
When m = 1, so that B reduces to a column vector b, the condition on
U in (4.5) becomes identical to that in (3.49). For this reason a system
described by (3.44), (3.45), and (3.46) is said to be in controllable canon-
ical form, since Theorem 3.4 can now be restated as:
THEOREM 4.2. A system in the form (3.48) can be transformed into
the canonical form (3.44) if and only if it is c.c.
Notice the important point that if the system model is taken to be in
the classical form (3.42), then by virtue of Theorem 4.2 this automatic-
ally assumes that the system is c.c. Thus the concept of controllability
could not be appreciated until linear systems in vector—matrix form
were studied, and this is one of the reasons why controllability consider-
ations have only recently been taken into account. Note also that
another remark on Theorem 3.4 can now be interpreted as stating that if
A is derogatory, (3.48) cannot be c.c. for any vector b.
Theorem 4.1 gives a criterion for determining whether a constant
linear system is c.c., but gives no help in determining a control vector
which will carry out a required alteration of states. We now give an ex-
plicit expression for such a vector for both constant and time varying
systems.
THEOREM 4.3. The system S, is c.c. if and only if the m x n symm-
etric controllability matrix
Proof.
=fa Wal,
2dr 0
>0, (4.14)
where 0(1, fo) = B! (1) ! (to, rT)a, so that U(to, ¢,) is positive semi-
definite. Suppose there exists some & # 0, such that @'U(to, t;) @=0.
Eqn (4.14) with @= 6 when a = a then implies
fsBl6tee
||, dr=0
which in turn implies, using the properties of norms (see Section 2.7)
that 0 (7, to) =0,t, <7<t,. However, by assumption S, is c.c. so there
exists a control v(t), say, making x(t, ) = 0 if x(to) = @.Hence from (4.3)
it follows that
@=- [- ®(to,r)B(r)v(r)
dr.
a
Therefore
=0,
which contradicts the assumption that @#0. Hence U(to, ft;) is positive
definite and is therefore nonsingular.
Since Theorem 4.3 shows that controllability of S, is independent of the
matrix C, we shall often refer to the controllability of the pair [A, B]
90 Linear control systems
instead of that of S,. The control function (4.13) which transfers the
system from (Xo, fo) to (xf, t,) requires calculation of the transition
matrix and the controllability matrix (4.12). This is not too difficult for
constant linear systems, although rather tedious, and moderates the
objection raised at the end of Example 4.3, Of course there will in general
be many other suitable control vectors which achieve the same result,
but the expression (4.13) has an additional interesting property:
THEOREM4.4 If a(t) is any other control taking (xo, fo) to (xg, f;)
then
ty
O= ['* (0, A
7)B(r) [a(r) -u(r)] dr:
[xo —
P(to, ty xe)" [U-* (to, t1)]
and useof (4.13)gives
ie u?(7)[u(r)—aa(r)]
dr =. (4.15)
Therefore
[ ty @ay
~ @aar=
~ res [aiea
[" +a A ar
2uTa)
0
pls A
= J, Cate ele] dr,
using (4.15), so
providedu # u, as required.
Linear control systems 91
| eAIlu(r)112
dr=f ie (u;?+++*+Um?)
dr,
Llanes t 0
over the set of all controls which transfer (xo, fo) to (xf, f, ), and this
integral can be thought of as a measure of control ‘energy’ involved.
If the system is not c.c., it would be misleading to call it ‘uncontrol-
lable’, since the implication of our definition is that for a non-c.c. system
there are only certain final states which cannot be achieved by any
choice of control. We can however modify the argument used in Theorem
4.3 to show how such final states can be attained when §, is not c.c.
THEOREM 4.5. If, for a given xf, there exists a constant column n-
vector y such that
x(t;)=@(t,,
to)[x0=|i: P(to,T)
B(1)B'
(1)®(to,
1 7)vdr|
=@(t,, 0) [xo— U(to,t1)¥] by (4.12)
=@(t,,fo) P(to,ti) X¢ by (4.16)
= xf, as required.
When §, is c.c., Uis nonsingular and the expression for u(t) in Theorem
4.5 reduces to (4.13). It can also be shown (Brockett 1970, p.77) that
the converse of this theorem is true, namely that only states x¢ for which
(4.16) holds can be reached.
92 Linear control systems
THEOREM 4.7. If ®(f, to) is the state transition matrix for S,; then
P(t) ® (t, to) P ' (to) = P(t, fo) is the state transition matrix for S).
Proof. From Section 3.5. we recall that ®(f, to) is the unique matrix
satisfying
and is nonsingular. Clearly (to, to) =/, and differentiation of (4.17) and
use of (4.1) gives
dx/dt = Px + Px
= (P + PA)x + PBu
= (P+ PA)P"! x + PBu (4.18)
To establish the theorem we must show that © is the transition matrix for
(4.18), ice.
Proof. From (4:18) we have that the system matrices for S, are
x es a & &
U(to,t1) = iP P(to, T)B(1)B' (1) ®1(to, T)dt
=[ Po)®60,
PPB OBDP(0)
x(P71(1))T"(to,7)P"(
= P(to) U(to,t1)P' (to) (4.20)
using (4.19) and Theorem 4.7. Thus U is nonsingular since the matrices
U and P(t) in (4.20) each have rank n.
When A, B, and C are time invariant P is also time invariant, and (4.17)
is the usual definition of equivalence transformation in matrix theory.
The following important result on system decomposition then holds:
THEOREM 4.9. When §, is time invariant then if U in (4.5) has rank
n, <n there exists a system algebraically equivalent to S; having the
form
d xO) Ay Az xO) B,
aH = + u (4.21)
x2) Orley x2) 0
y=[C, C,|x
Exercise 4.1.
Verify that the system in Example 1.1 is c.c.
Exercise 4.2.
_ Re-interpret the result of Exercise 3.31 in the light of Theorem 4.2.
94 Linear control systems
Exercise 4.3.
Two pendulums are connected by a spring as shown.
Spring constant =k
mg mg
FIG. 4.1
The masses of the rods, which have unit lengths, can be neglected. Equal
and opposite control forces u(t) are applied to the particles, which have
mass m. Write down the equations of motion for small oscillations so that
6? and higher powers can be neglected. Take x; =0, +6,xX2 =O; —02,
X3 =X ,,Xq4=X, as state variables and hence deduce that the system is
not c.c.
Notice that if the control forces were not equal the system would be
Cc
Exercise 4.4. f
Consider again the problem of two connected masses moving on a smooth
horizontal plane, described in Exercise 1.2. Suppose now that in addition
to the springs the masses are joined by dampers as shown, the forces as
usual being proportional to velocity with coefficients d, and d>.
Linear control systems 95
Take the same state variables as before and derive the state equations in
the form (4.4). If m, =m, =1,d,; =d, =1andk, = determine under
what conditions the system is c.c.
Exercise 4.5.
A platform is supported by springs and dampers as shown, it being assumed
that the forces they produce act at the end points P and Q, and that x,
and x2 are the displacements of these points from equilibrium.
4u
FIG. 4.3
Xy ull 0 X41 1
X4 On? X2 8
Verify that the system is c.c. If the ends of the platform are each given
an initial displacement of 10 units, find using (4.13) a control function
which returns the system to equilibrium at ¢ = 1.
Exercise 4.6.
If in the system (3.48) A has repeated characteristic roots but neverthe-
- less the Jordan form of A is a diagonal matrix, show that (3 48) is not c.c.
96 Linear control systems
Exercise 4.7.
If the pair [A, B] is c.c. show that [A + BK, B] is also c.c. for any matrix
K having appropriate dimensions. (Hint: do not use Theorem 4.1).
Exercise 4.8.
Using the matrix A in Exercise 3.1(a), find for what vectors b the system
(3.48) is not c.c. Using Laplace transforms investigate what happens to
X(s) in these cases.
Exercise 4.9.
By taking A to be the companion matrix
0 0 math
0 Sols
0 mln
Exercise 4.10.
Prove that U(to, t, ) defined in (4.12) satisfies the matrix differential
equation
U(t,t1) = A(0)U(t,
1) +U(t,t))AT(1)-B(OBT
(1),Uy, t1)=.
Exercise 4.11.
In the preceding exercise let A and B be time invariant, and put
W(t, t, ) =U(t, t; )—Up where the constant matrix Up satisfies
AU + UpA! = BB".
U(t,t,) =Up—
exp[A(t—11)]
Uoexp[AT(t-1t,)].
Linear control systems 97
Exercise 4.12.
If
] 2 ma 0)
x=) 0 1 Oe ae Ou
1 —4 3 1
find a basis for the set of vectors which can be reached from the origin.
Exercise 4.13.
Show that the result in Theorem 4.4 still holds if the measure of control
energy is
[Put
@R@uGyar
0
4.2. Observability.
X, =a,xX, +byu
X_ =d,xX2, +b2u
yHx,.
Vitosts)=
fOr, to)CTCO)®(r,
fo)dr, (4.22
where ® is defined in (3.27), is nonsingular.
Proof.
A aT T
J 87G, t0)C Op dt = V(t0, 110,
SOS 1SC10,,
eo =i if[C(r)
®(r,toa]T[C(r)®(r,
0
toJor]
dr
20
v=| - (4.26)
Com
has rank n.
2 2
which has rank 2. Thus the system is c.o., that is, it is possible to deter-
mine the temperature of the oven interior from a knowledge ofu(t) and
x,(t).
It is worth noting that in the scalar output case (i.e. = 1)ifu=0
y(t) is known in the form
n
ESviexp(Ajt),
assuming that
i=]
all the characteristic roots A; of A are distinct,
then x9
can be obtained more easily than by using (4.24). For suppose that
to = 0 and consider the solution of x = Ax in the spectral form (3.5),
namely
We have
THEOREM 4.13. A system in the form (3.60) and (3.61) can be trans-
formed into the canonical form (3.62) if and only if it is c.o.
For this reason a system described by (3.62), (3.63), and (3.64) is
‘often said to be in observable canonical form.
Again by duality the result corresponding to Theorem 4.9 is:
THEOREM 4.14. When S, is time invariant then if V in (4.26) has
rank n, <n there exists a system algebraically equivalent to S, having
the form
aes Aas 00 x1) B
eat = 4 u (4.27)
Re (2) A, A3 x) B,
PC R™
where x‘!),x?) have orders n; andn—n, respectively and [4;, C;]
is C.0.
As in Example 4.4, we see that in (4.27) y is completely unaffected
by x?) which is therefore said to be unobservable, and the state space
has been divided into two parts with respect to observability.
We close this section with a decomposition result which effectively
combines together Theorems 4.9 and 4.14 to show that a linear time-
invariant system can be split up into four mutually exclusive parts,
_ respectively (1) c.c. but unobservable (2) c.c. and c.o. (3) uncontrollable
102 Linear control systems
and unobservable (4) c.o. but uncontrollable (see Zadeh and Desoer,
19633 p27505);
THEOREM 4.15. When S, is time-invariant it is algebraically equivalent
to
x) All Ate Ais. gai xi B}
d x2) a 0 A? OQ Aw x2) B?
aft ,@)| 19 96 Byala eae %
x4) OiAiulOs
oof:edBebbaa
y= Cx) + C44
Exercise 4.14.
If the output in the spring-mass system described in Exercise 1.2 is taken
to be y =x, is the system c.0.?
Exercise 4.15.
Consider again the rabbit—fox environment problem described in
Exercise 1.4. If it is possible to count only the total number of animals,
can the individual numbers of rabbits and foxes be determined?
Exercise 4.16.
In the circuit shown in Fig. 4.4 take the state variables to be the
voltage x, across the capacitor and the current x, through the inductor.
The input voltage u is the control variable and the current y can be
regarded as the output. Derive the equations describing the system, and
show that if R,;R, C=L it is not c.c. and not c.o.
FIG. 44
Linear control systems 103
Exercise 4.17.
Using the matrix in Exercise 3.1 (@),find for what vectors c the pair
[A, c] is not c.o. By using the solution of (3.1) obtained, investigate
what happens to y(t) in these cases.
Exercise 4.18.
Bearing in mind the result of the preceding exercise, does it follow that
if some of the exponential terms present in x(¢) are missing from y(t)
then the system is not c.o.?
Exercise 4.19.
For the system xX= Ax with A matrix as in Exercise 3.1 (@’)and y = [1,2]x
find x (0) if y(t) =—20 exp (—3r)+ 21 exp (-22).
Exercise 4.20.
Suppose that the pair [A, c] is c.o., where c is a row n-vector, and that
all the characteristic roots of A are distinct. If the right characteristic
vectors of A are W,,..., W, Show that cw; #0,i=1,2,...,n.
"Exercise4.21.
If in the system described by (3.60) and (3.61) A has repeated character-
istic roots, but the Jordan form of A is a diagonal matrix, show that the
system is not c.o. (compare with Exercise 4.6).
Exercise 4.22.
Prove that V(to, t, ) defined in (4.22) satisfies the differential equation
V(t, 1) =-AT(*)V(t,t,) V(t, t1AQ) CT(DCO, Vi, t1) = 0.
Compare with Exercise 4.10.
Exercise 4.23.
Prove directly the necessity of the condition in Theorem 4.12. (Hint:
assume rank V <n and obtain a contradiction).
Exercise 4.24.
Show that if the initial state of a constant system which is not c.o.
satisfies Vx(0) = 0, where V is defined in (4.26), then y(t) = 0 for all
- t2>0. (Hint: it can be assumed that the system is in the form (4.27).)
104 Linear control systems
Exercise 4.25.
Obtain the result corresponding to Theorem 4.8 by expressing the observ-
ability matrix V(t, t,) in terms of V(to, f; ).
0 0 : : Ky,
] 0 . : “Kye
A= : ; : ; 6 (4.28)
0 0 : 1 —ky
then the basic Theorem 4.1 on constant systems can be stated in the
following form:
THEOREM4.16. The pair [A, B] is c.c. if and only if the set of poly-
nomials
k(x) =det (MU
-A)=N? +k” 1 +e +e thy
and
Pj(A)=by? +byi id” 2 +92 ++b,,,15
where B = [b;;], is relatively prime. Furthermore, the columns of the
controllability matrix in (4.5) are a re-ordering of the columns of
If we write
qi(s)
Sy) Sty
(si —=A)| B aE@ ed ,
qn(s)
106 Linear control systems
then
qilk
Also
1 s+4 = b,
heyy2
Ce ae Leet
Tan? Pa
s? +55+6
(s +4)b, —b2
1
oeCaisbak’
| Geep
The pair [A, b] is not c.c. if bz = 2b, or by = b,, in which cases there
are common factors s + 2 and s + 3 respectively in the vector transter
function.
Exercise
4.26.UsetheCayley—Hamilton
theorem
toshowthatif
D(A)=bod"+byNt te ee+d,
Linear control systems 107
then
DCD)1640. Tare Gl
where
Aisgiven
by(4.28)and
B=[bn—Dokn, bn1 —bokn-1,-..,51—bok,]".
Exercise 4.27.
Using the notation of Exercises 2.12 and 2.13 deduce that the pair [C, r]
is c.o. if and only if k(A) and P(A) are relatively prime.
Exercise 4.28.
If k (A) denotes the derivative of k (A) with respect to A, and A is defined
in (4.28), deduce that k’ (A) is nonsingular if and only if k (A) has no
repeated zeros.
u= Kx (4,32)
where K is a constant m x n feedback (or gain) matrix. The resulting
closed loop system obtained by substituting (4.32) into (4.31) is
%=(A+BK)x. (4.33)
Assume that as is usual in applications A and B are real, and let A, =
{6,, D5 eee On} be an arbitrary set of n complex numbers such that
any which are not purely real occur in conjugate pairs.
THEOREM 4.17. If the pair [A, B] is c.c. then there exists a real matrix
K such that the characteristic roots of A + BK are the set Ay.
In view of the development in Section 3.2 it follows that the solution
- of (4.33) depends on the characteristic roots of A+ BK, so provided
108 Linear control systems
(4.31) is c.c. the theorem tells us that using linear feedback it is possible
to exert a considerable influence on the time behaviour of the closed loop
system by suitably choosing the 6;. A further comment on this point will
be made at the end of the section. However, it should be noted that it is
not, in general, possible to choose K in order to give A + BK an arbitrary
Jordan form..
The theorem gives us a new and most interesting insight into the
meaning of linear feedback in state space terms. For example, if a single-
input single-output system is in the classical form (3.42), then application
of Theorem 4.17 to the equivalent matrix representation (3.44) shows
that in general u will be a linear combination of all the phase variables,
that is of z,z“),..., z—), not just a multiple of the output z alone.
If (4.31) is not c.c. then it follows from Theorem 4.9 that only n, of the
characteristic roots of the closed loop system can be arbitrarily chosen
as members of a set Ay, , since A3 in (4.21) is unaffected by the feed-
back.
The theorem is very recent, the first general proof being given by
Wonham in 1967. We shall present a somewhat different argument which
is nevertheless rather lengthy, so we first consider the case when there is
only a single control variable, and give a proof which includes a method
of constructing K.
Proof when m = 1. Since (4.31) is c.c. it follows from Theorems 4.2 and
3.4 that there exists a nonsingular transformation w = 7x such that (4.31)
is transformed into w = Cw + du where C and d are given by (3.45) and
(3.46) respectively. The feedback control u = kw where
KilKnikn=ieetae kal
produces the closed loop matrix C + dk which has the same companion
form as C but with last row —[Yn, Yn—1,---, 71], where
it follows that the desired matrix is K= KT, the x; being given by (4.34).
In this equation the k; are the coefficients in the characteristic poly-
nomial of A, i.e.
det (MW
-A)=A" +ky tl te + +tky, (4.36)
Linear control systems 109
Beat n
nctn LT0-81). (4.37)
The realness of K follows from that
l=of x and T.
Notice that (4.35) shows that in the single control variable case the
closed loop matrix A + b«T is similar to acompanion form matrix, and
so in nonderogatory (see Section 2.5). This provides an illustration of
our earlier remark that the closed loop matrix cannot be made to take
an arbitrary Jordan form. Notice also that if in the preceding argument
the canonical form involving C, d is replaced by the observable canonical
form in Theorem 3.5, we obtain:
COROLLARY 4.2. If the pair [A, c] is c.o., where c is a real row
n-vector, then there exists a real column n-vector & such that the charac-
teristic roots of A + &c are the set A,.
This result can also be deduced from Theorem 4.17 using the duality
Theorem 4.11.
N?- 3+ 14=0
which has roots 3/2 + i\/(47)/2. Suppose we wish the characteristic
roots of the closed loop system to be —1 and —2, so that the corres-
ponding polynomial is \* + 3d + 2. From (4.34), (4.36), and (4.37)
we have
hg Jeo
Ky,=k. -Y¥2=14-2=12.
K=kT
= % [12,-6] [-1
posters 1
=— [15/4, 9/4].
110 Linear control systems
a
al
A+bK=% eee
|
El
does have the desired roots.
Before proving Theorem 4.17 for m > 1 we need a preliminary result.
LEMMA 4.1. If the pair [A, B] is c.c. and the columns of B, assumed
non-zero, are b,,b2,..., 5), then there exist real matrices K;,i= 1, 2,
..- ,@t, such that the pairs |A + BK;, b;) are c.c-
M=[b,,4bi,...;4% ~01,0,,405,....42 2
where r; is the smallest integer such that A”! ; is linearly dependent on
all the preceding vectors, the process continuing until 7 columns of U
are taken. Define an m x n matrix N having its 7; th column equal to
€2, the second column of Jy, its (ry; +72 )th column equal to e3, its
(r; +72 +73 )th column equal to eg and so on, all its other columns
being zero. It is then not difficult to show that the desired matrix in
the statement of the Lemma is K, = NM~'. This is established by comp-
aring terms on both sides of the expression K,M = WN,from which it
follows that
b,, (4+BK,)b,,(A+BK,)°b;i,...,(4+BKiy”Oy
are linearly independent, so by Theorem 4.1 the pair [A +BK,, b, ]
ISiCuCe
It is worth remarking that the matrix M defined above can be used
as a transformation giving a canonical form for (4.31) with m > 1 (see
Chen 1970, Chapter 7 where fuller details of the proof of Lemma 4.1
can also be found).
where b, is the first column ofB. Since the pair [A + BK, ,},] isc.c.
it now follows from the proof of the theorem when m = 1 that k can
be chosen so that the characteristic roots of A +BK, + b,k are the set
A, so the desired feedback control is indeed (4.38).
If y = Cx is the output vector in (4.2) then again by duality we can
immediately deduce from Theorem 4.17:
COROLLARY 4.3. If the pair [A, C] is c.o. then there exists a real
matrix L such that the characteristic roots of A + LC are the set Ay.
It is interesting to relate the feedback theorem to the transfer function
for the scalar input and output case defined in (1.14), namely
Base PRS is Phe kB
g(s)2
st +k? tee +k,
It is easy to verify (see Exercise 4.39) that
g(s)=r(sI-—C)"'d (4.39)
where G(s) is the r x m open loop transfer function matrix. The closed
loop system considered is of the form shown in Fig. 4.5.
FIG. 4.5
This is the multivariable version of Fig. 1.8 and the equation correspond-
ing to (1.15) is
u(t)=—y(t)
=—Cx(t)
so that the closed loop system is
x =(A-BOC)x.
An interesting relationship between the two systems is the following:
THEOREM 4.18. The ratio of the characteristic equations of the closed
Linear control systems 113
loop and open loop systems is equal to the determinant of the return
difference matrix, i.e.
eee BO) _ a
det (sI, -A +B [Im + G(s)] (4.41)
Proof.
det(s!—A+BC)=det[(s!—
A){I+(sJ]-—
A)"'BC}]
=det(sJ —A)det [J+ (sJ—A)7'BC]
=det(sJ—A)det [J+C(sJ—A)"'B],
= det (s/— A) det [J+ G(s)],
the penultimate step following from the fact that for any two matrices
X and Y having dimensions n x m and m x n respectively then
det U, +XY) = det Um + YX) (4.42)
(see Exercise 2.3).
The proof of Theorem 4.17 for m > 1 given earlier does not provide
a practical way of constructing the feedback matrix K in (4.32). However,
we can use (4.42) to derive a simple method due to MacFarlane:
THEOREM 4.19. Let all the characteristic roots \; of A be distinct and
let W=[w1, W2,..., Wy], where w; is a characteristic vector correspond-
ing to A;. With linear feedback u = —Kx, suppose that the characteristic
roots of A and of A —BK are ordered so that those of A—BK are to be
Mi, U25-++,Mp,Apt+1,---5An (PpSn). Then provided [A,B] isc.c.a
suitable matrix is
K =few (4.43)
whereWconsistsof thefirst prowsof W~!,and
&= [01/B1,&2/B2,.-.,%/Bp), (4.44)
a =T1 Ov-4)/TT Ov-¥).2 > 1, (4.45)
a ii
6 = [B1,B2,--.,8p]" = WBF, (4.46)
f being any column m-vector such that all 6; # 0.
114 Linear control systems
Proof. Since only the first p characteristic roots of the closed and open
loop systems differ, eqn (4.41) gives
det[1+K(sl—A)"'B]
=II (s—ui)/ia (s=)j). (4.47)
i=1 i=l
Furthermoresince A= WAgW~',whereAg = diag [\1,..., An], (see
(2.31))wehave
where the first and last steps follow from (4.42). Using the definition of
W the right side of (4.48) becomes
1+g(s] -—Aq)”!WBf
=1+g(s!—Ag) 1B (4.49)
Dp
=1+ >) a;/(s —4), (4.51)
i=
where the a; given in (4.45) are obtained by expansion of (4.50) into
the partial fractions (4.51) (note that when p = 1, a, =A, —4y). Clearly
so the expression (4.44) for the elements of g follows from (4.51). Since
the system is c.c., WB has no zero rows (see Exercise 4.37) and this
ensures the existence of a suitable /.
The method can be extended to the case when A has repeated roots.
Note that the relative simplicity of the expression for K is due to the
fact that the matrix on the right in (4.43) has rank one.
Linear control systems 115
Example 4.8. Consider the matrix A given in Example 3.1, and let
b= [2,1] 1. Wehave \y =— 1, A. =—2 and
1 1 2 1
= ; whe :
“all Ry eee) alee el
By 2s 1 2 Shi
im i= :
Bo am| =] 1 i 3f;
&= 16/52/31
ia | Sit*
B, Seely
=fr
so that f; = 1,f, =0 gives K with first row identical to (4.52) and second
116 Linear control systems
K= (2. 1h
-1
3 1
-3 -1
and again the reader can verify that A —BK has the roots —3 and —4.
Methods for controlling systems by prespecifying the closed loop
system characteristic roots are often called modal control techniques,
the name mode being given to the terms exp(A;t)w; in the expression
(3.5) for the solution of a linear system. Much effort has been put into
this approach (Porter and Crossley 1972) but there are anumber of
disadvantages which arise in practical applications. For example, there
is no means of directly correlating the time behaviour of the closed
loop system with the specified values of the poles since, as we have men-
tioned, there is no control over the closed loop zeros (see MacFarlane
(1972) for further discussion).
Exercise 4.29.
Using the matrices in Exercise 3.30(a), use the method of Theorem 4.17
to find a 1 x 2 matrix K such that the closed loop system has charac-
teristic roots —-4and —S.
Exercise 4.30.
Repeat the procedure of the preceding exercise using the matrices in
Exercise 3.30(b) so as to make the closed loop system characteristic
roots —Ix=1"= 21.
Exercise 4.31.
’ Verify that the method of Theorem 4.17 when applied to Example 4.8
produces the same matrix K.
Linear control systems 117
Exercise 4.32.
If
0 1 0 1 0
A=|0 0 iad pe B= | 0 1
6 = 11 6 1 1
find using Theorem 4.19 a suitable matrix K which makes the character-
istic roots of A —BK equal to 1, 1, 3.
Exercise 4.33.
Consider yet again the rabbit—fox environment problem of Exercise 1.4.
It was seen in Exercise 1.7 that if a, —a@4 >0 in eqns (1.6) then in general
the animal populations increase without limit ast > °°. Suppose for
simplicity that the values of the constants are such that the system
equations are
xy = 2x, = Shey,X Ph a ae
It is decided to control the environment by introducing linear feedback
_in the form of a disease which is fatal to rabbits but does not affect
foxes, thereby reducing the rate of growth of the rabbit population by
an amount kx,. Find the smallest value of kKwhich prevents the “‘popula-
tion explosion’ from occurring (assuming arbitrary initial population
sizes).
Exercise 4.34.
Consider
X=Ax+bu, u=cx
where b and ¢ are respectively column and row n-vectors. Let b be a
right characteristic vector of A corresponding to a root A. Show that
b and \ + cb are corresponding characteristic vector and root of the
closed loop system.
If on the contrary b is arbitrary but c is a left characteristic vector
of A corresponding to A, show that c and ) + cb are corresponding
characteristic vector and root of the closed loop system.
Exercise 4.35.
In the preceding exercise let b and c be respectively right and left
characteristic vectors of A corresponding to the root A. Assuming that
118 Linear control systems
all the characteristic roots of A are distinct, show that the other roots
and vectors of the closed loop system are just the remaining roots and
vectors of A.
Exercise 4.36.
Let all the characteristic roots of A be distinct and let W be the matrix
defined in Theorem 4.19. By applying the transformation x = WE to
(4.31) deduce that wu;can influence £; if and only if pj; # 0, where the
n
pjj are defined by b; = » pjiw;, bj denoting the ith column ofB. Note
7-1
that as in Section 3.1 it follows that pj; = v;b; where v; are left character-
istic vectors of A.
If QO;
= bj, Abj,A7b;,..., A"—By]
it can be verified that Q; =WP;M’ where M is the Vandermonde matrix
defined in Exercise 2.15 and P; = diag [p1;, Daj, . . - ,Pu]. Hence deduce
that rank Q; is equal to the number of modes which can be controlled
independently by the input w;.
Exercise 4.37. 4,
Using the notation in Theorem 4.19, deduce that if WB has a zero row
than the associated controllability matrix (4.5) has rank less than n.
y=, (4.53)
then linear output feedback
u=Ky = KCx (4.54)
is applied, but is no longer possible to preassign all the closed-loop
characteristic roots. However, if C has rank r, B has rank m and the
system is c.c. and c.o, then it has recently been shown (Sridhar and
Lindorff 1973) using an argument similar to that in the proof of
Theorem 4.19 that a matrix K exists such that at least s =max (m, r)
Linear control systems 119
= Goeee)
i= ail.C)2)
SO
dx/dt=%—(A+LC)(x-x)
=Ax+Bu-(A+LC)(x-x)
=tA+tLO)x hy +Bu, (4.56)
Thus (4.56) represents a mathematical model for the desired state
observer, having inputs y and wuand output x.
Example 4.10. Consider the case of scalar control variable and output,
let the matrices A and b be as in Example 3.8, and take C = [2,3]. To
find the 2 x 1 matrix L we use a procedure similar to that given in the
proof of Theorem 4.17 for m = 1, but commencing with the observable
canonical form in Theorem 3.5. The transformation is vy=P"! x, where
from the result of Exercise 3.36 we have
ey 1 CG
pis ;
Law CA
120 Linear control systems
where
This gives
8 -9
po =
phiae
0), a14
E=P"'!AP= ;
1 3
f=CP=[0,1].
Suppose that the characteristic roots of the observer system matrix
A + LC are to be —3 and —4, so that its characteristic equation is
? + 7d + 12. Since
A+LC=P(E+P' LAP!
Hence L = [-2, —2]T the matrix required for constructing the observer
system (4.56).
An important point which is now demonstrated is that although the
feedback matrix K was obtained in Section 4.4 with respect to the actual
state x, use of linear feedback u = Kx instead of u = Kx still results in the
closed loop system having the desired set of characteristic roots. To see
this, put u = Kx in (4.31) and (4.56) and combine these two equations in
the form
x A BK G
dt ox -LC (A+LC+BK) x
Linear control systems 12
Exercise 4.38.
Using the matrices A and c in Exercise 3.32, find a matrix L such that
A +Lc has characteristic roots —3, —4, and —5.
G(s) = C(sl—A)7'B,
and the system equations will then be (4.31) and (4.53). The triple
{A, BSChis termed a realization of G(s) of order n, and is not, of course,
unique. Amongst all such realizations some will include matrices A having
least dimensions —these are called minimal realizations, since the corres-
122 Linear control systems
(s1]—AY=Adj(sI—
A)/det(sf—A)
has the degree of the numerator less than that of the denominator it
follows that C(s] -A)"'B > Oass > ©, and we shall assume that any
given G(s) also has this property, G(s) then being termed proper.
As Kae i
(8)sSj= s* —5s+6°
Bee wa U
A= b= EN).
6 5 1
2enn0 1
A= es c= [-11,13]. (4.59)
Cables 1
In fact both these realizations are minimal, and there is in consequence
a simple relationship between them, as will be shown later.
We begin our development with a generalization of the result of (4.39),
an example of which has just been given. This enables us to write down a
simple realization for a transfer function matrix, although it will not in
general be minimal.
THEOREM4.20. Let
be the monic least common denominator of all the elements gj; (s) of
G(s), and let
2(s)G(s)=541 Gp+897
7G, +¢ °° +Gq-1, (4.61)
Linear control systems 123
0 ie ei 0
0 0 If,
A= wo= q
Im eee (462)
Bqlm Sq-tlm + Pcs ad 0
Im
CamereG, >,..-4 Go].
B = eq ®lm, (4.64)
where eg denotes the last column of J/g, so (4.63) and (4.64) give
1
S
s7 Q/,/2(s), (4.65)
57-1
124 Linear control systems
using the expression given in Exercise 4.39 for the last column of
(slg -F)" = Adj (slg —F)/g(s). Finally, combining (4.65) and the
expression for C in (4.62) we have
C(sI- A) "B= [G,_,,G,_>)° a ,Go] Im /g(s)
Sl,
Rseae be
= G(s),
by virtue of (4.61), showing that{A, B, Chis a realization of G(s). f
That [A, B] is c.c. can be seen from the fact that the controllability
matrix (4.5) is
0 00 ee ee
0 0 . ° ° a¢ L€ “ome
C(sl —
A) B=C(sI—A)"B.
We can now prove the central result of this section, which links
together the three basic concepts of controllability, observability, and
realization.
C(sl-A) 1B=C(sI-A)'B
it follows(seeExercise4.45) that
CAB= CA Bei=
0, 122,.13,» (4.67)
126 Linear control systems
G
CA
VU=|- [B, AB....,A" "sie
GA" —1
CB CABS WR CA™"B
C
CA
= (B, AB,...,A”!B], using (4.67)
CA nti
= V,U,, Say. (4.68)
By assumption V and U both have rank n, so the matrix V;U, also has
rank n. However, the dimensions of V, and U, are respectively r,;n x n
and n x m,n, where r, and m, are positive integers, so that the rank of
V,U, cannot be greater than n. That is, n <A”, so there can be no real-
ization of G(s) having order less than n.
Necessity. We show that if the pair [A, B] is not c.c. then there exists
a realization of G(s) having order less than n. The corresponding part of
the proof involving observability follows by duality.
Let the rank of U in (4.5) be ny <nand let u,,u2,..., Uy, be any
set of n; linearly independent columns of U. Consider the transformation —
& = Px with the n x n matrix P defined by
A =PAP™', by (4.66),
= A, Az
0 A3
where A, ism, xXny. Similarly, since u,,...,uU,. also forms a basis for
the columns of B we have from (4.66) and (4.69)
By
=VRC 0 1/R,C
A= p= ,c= |-l Ryan
0 Seuss 1/L
(s) =Se(sI-—
A)“1b+1/Ry}
Us)
andhence
g(s) M(RTC=E)s +(R; —R2) + m9:
(Ri Cs+R,) (Ls+R) R,
By the result of Exercise 4.16 it follows, using the theorem just proved,
that the above triple {A, b, c} is a minimal realization for g(s) if and only
if L #R,R,C In fact it is easy to verify that when L = R,R.C there
is a common factor between numerator and denominator of g(s).
Of course for any given G(s) there are an infinite number of minimal
realizations satisfying the conditions of Theorem 4.23. However, we now
show that the relationship between any two minimal realizations is just
that of algebraic equivalence:
VU=VU, (4.71)
pu=U (4.73)
P= tu ™(uuT)", (4.74)
K; = L;M; , (4.76)
130 Linear control systems
Sil; 0 ee
M,
A= Sol,
ie ee he (4.77)
0) . a
Spry | Dp
C= ese At Lp].
C(sl —ANyye =i diag Uy, [(s wo ; I, Ms= Se, tee Try l(s —Sp)) B
= ZL Mj |(s —8)
(Bee e va An Bl ae (4.79) |
Mp SpMp 2
SpMp « Sp
git Mp
where n=r, tr, + +++ +rp. Using the facts that M; has rankr;
and all the s; are distinct it follows that the matrix in (4.79) has full rank
(see Exercise 4.51). The observability condition is proved similarly.
Linear control systems ti
s-|my,
| c=[—11/m,,
13/m,]
mM
can be used instead, still giving a minimal realization for arbitrary nonzero
values of m, and m).
Notice that (4.75) can be used to evaluate the order of minimal
_realizations of G(s) without actually determining a realization, since the
p
order is the dimension of A in (4.77), namely > r;. However, calculation
=|
of the 7; involves the same disadvantage as using Theorem 4.25 in general,
namely the need to calculate the s;. An alternative method of finding the
least order of G(s) which does not rely on any knowledge of the s; is
obtained by arguing as follows. The realization in (4.62) is c.c. but not
in general c.o. By Theorem 4.14, (4.62) is therefore algebraically equi-
valent to
A, O B,
5) > [C, p)0] (4.80)
Az A; By
Exercise 4.39.
If C, d, and k are as defined in (3.45), (3.46), and (3.47), and
G(s) =Bos” 1 +B,s?~ 2 +++ ++B, —4, verify that a realization of
B(s)/k(s) is {C d, r} where r= [8,—1;B8n—2;---,>6o]. (Hint: By con-
sidering the product (s/ —C)Adj(s/ —C) show that the last column of
Adj(sI —C)is [1,s,s?,...,8”~4]").
Show that the realization is minimal if 6(s) and k(s) are relatively
prime. (Use the result of Exercise 4.27).
Exercise 4.40.
Deduce from the result of the preceding exercise that a realization of
B(s)/k(s) is ew rie d‘} and that this is c.o. but not c.c. if B(s) and k(s)
are not relatively prime.
Exercise 4.41.
Let A be an arbitrary n x n matrix with characteristic polynomial
q(s), and let
c(sl —
A)" 'b = t(s)/q(s) =8(s)
where cis 1 x nandbisn x 1. Prove that {4,b,chis a minimal real-
ization of g(s) if and only if t(s) and q(s) are relatively prime.
Exercise 4.42.
Using the notation in Theorem 4.16, let
Px(S)
Eas:
G(s) = KS)
Pm(S)
and assume that the polynomials k(s), p;,..., Dm are relatively prime.
Generalize the result of Exercise 4.39 by verifying that a minimal real-
Linear control systems 133
bim i ; bam
Exercise 4.43.
Using Theorem 4.25, write down a minimal realization for a scalar
transfer function whose denominator has only simple zeros.
Exercise 4.44,
Let G(s) and g(s) be as defined in (4.60) and (4.61). Deduce that
{A™,B, C}, where A is defined in (4.62),
Go-1
Exercise 4.45.
Using (4.66) prove Theorem 4.22, and hence deduce that
C{exp(At)}B
=C.fexp(A
t)B.
Exercise4.46.
UseTheorem4.25 to obtain a minimalrealizationof
(s? +6) (s? +5 +4)
g(s)
(Qs* 7s 2) (Ss? 55—2)
where g(s) =s° + 2s? —s—2.
134 Linear control systems
Exercise 4.47.
Show that the order of a minimal realization of
(s+2) 2s +2)
s? +3s+2
i (srtih)
is three. (Notice the fallacy of assuming that the order is equal to the
degree of the common denominator).
Exercise4.48.
If{A,,B,,C,}and {A,,Bz,C>}arerealizations
of G,(s)andG,(s),
showthat
A, B,C, 0
A= , B= , C=[C, 0]
0 A, B,
Exercise 4.49. .
Verify that in the second part of the proof of Theorem 4.23 the pair
[A,,B,] isc.c.
Exercise 4.50.
Show that the transfer function matrix of the system in Theorem 4.15
depends only on the c.c. and c.o. part. This result shows that the
uncontrollable and unobservable part of the state space representation
is lost when the transfer function matrix description is used.
Exercise 4.51.
In (4.79) let p =2,r,; =2,r2 = 2 for simplicity. By considering the
linear independence of the rows of the matrix on the right in (4.79)
show that it has full rank.
Exercise 4.52.
Verify that algebraic equivalence in (4.66) can be written as the trans-
formation
Linear control systems 135
Rearranging(4.82) gives
x(z) = (zl —A)"1Bu(z)
so the output y (kK)
= Cx(k) hasz-transform
p(z) = C(I —A) *Bu(z). (4.83)
Thus the transfer function matrix in (4.83) relating y(z) and u(z) has
exactly the same form as that for the continuous-time case. The theory
developed in Section 4.6 therefore carries over directly for (4.81)
virtually without modification, although in view of our remarks on the
case when A is singular, Theorem 4.23 should be interpreted as stating
that a realization is minimal if and only if the matrices (4.5) and (4.26)
have full rank. When linear feedback u(k) = Kx (k) is applied to (4.81)
the closed loop matrix is again A + BK, so the methods of Section 4.4
also apply to discrete-time systems.
Exercise 4.53.
It was seen in the buffalo population model in Exercise 1.13 that under
natural conditions the numbers would grow without limit. In fact due to
indiscriminate slaughter the number of animals was reduced from 60
million in 1830 to only 200 in 1887.
Suppose that a controlled slaughter policy had been adopted, killing
for food Sx + 1 adult females in year k + 1 so that (1.33) becomes
Fy 42=0-95Fx
41+0-12FE
- Sx 4.
If linear feedback Sx + 1 =pFx +1 is used, show that the total population
would remain constant even if 7 per cent of adult females were killed each
year. Also, find an expression for the number of adult females in the
steady state in this case.
What difference would it make if there were a year’s delay in com-
pleting the buffalo census, so that Sx + 1 =pFx?
As a footnote, a single buffalo carcase would provide enough meat
(250 kg) for at least 10 people for a whole year!
Exercise 4.54,
Show using (3.39) that when A is nonsingular, then if the control sequence
u(ko),...,U(N— 1) transfers x(Ko) =Xo to x(N) = xg, it also transfers
Xo —A”*° ~ “*x¢to the origin in the same time interval.
Linear control systems 137
Exercise 4.55.
Verify using (3.41) that the control sequence
We now return to systems described by eqns (4.1) and (4.2). Once again,
as in the case of the transition matrix for time varying systems discussed
in Section 3.5, we find that many of the properties established for time-
invariant systems in Section 4.6 still hold. However, as is to be expected,
' we cannot give analytical methods for calculation of realizations.
As in the constant case assume that x9 = 0. Then using (3.30) the
output is
¥())= Ci)x()
C(t)i b(t,1)B(r)u(r)
dr
K(t, ru(r) dr
is called the weighting pattern matrix. For a given K (t, 7) the realization
problem is now to find a triple {A(t), B(t), C(t) } such that (4.84) is
satisfied. The notion of minimality of a realization is the same as before,
that A should have least possible dimension. The relationship with the case
when A, B, and C are constant is obtained by noting that by (3.11) the
transition matrix is then
@(t, 7) = exp [A(t—-7)], (4.85)
138 Linear control systems
Necessity. We assume that the pair [A (t), B(t)] is not c.c. and show
that the realization R is then not minimal. A similar argument applies
if [A, C] is assumed not c.o. Suppose that A isn x n and that the
controllability matrix U(to, tf;) in (4.12) has rank p <n. Since, as was
shown in the proof of Theorem 4.3, U(to, t; ) is positive semidefinite,
there exists a nonsingular matrix T such that
jp iitero
TUP = (4.87)
On 10
(see Exercise 2.23). Consider the algebraic equivalence transformation
with P(t) in (4.17) taken to be TX (to )X~ '(t), where X(t) is defined by
Linear control systems 139
Sufficiency. We now assume that R is c.c. and c.o. and show that there
cannot exist a realization R, of K having order n, <n. First, by again
applying the result of Exercise 4.58 we can assume that such a realization
has the form R, ={0n, fora On (t)}. It similarly follows that under
the transformation P = X~'(t), R becomes R ={0;, , B(t), C(t)}and
remains c.c. and c.o. (see Theorem 4.8). Since
Wes
f‘ CTY)C\(t)dt,
W,=fBi BT(oar
and Vand U are the observabilityand controllabilitymatricesfor R
definedin (4.22) and (4.12) respectively.By assumptionV andU have
rank n, so VUalsohas rank n. However,W, and W, have dimensions
140 Linear control systems
Exercise 4.56.
Verify that in Theorem 4.26 suitable matrices are L (t) = C(t) ®(f, fo)
and M(t) = ®(to, T)B(r), where ®(f, to) is defined in (3.27).
Exercise 4.57.
Show, using Theorem 4.7, that if two realizations are algebraically
equivalent then the corresponding weighting pattern matrices are identical
Exercise4.58. |
Showthat if P(t) in (4.17)is taken to be EX '(t), whereX(t) is defined
in (3.26)and £ is a constantnonsingularmatrix,then{A(¢),B(t), C(t)}
istransformedinto{0,P(t)B(t),C(t)P™(t)}.
Linear control systems 141
Exercise4.59.
Showthat thesystem
X=Ax + {exp(Dt)}Bu,y =C{exp(—Dt)}x
where A, B, C, and D are constant matrices and AD = DA, possesses a
constant realization.
5 Stability
5.1. Definitions
x =f, a (5.1)
where as before x(t) is the state vector and f is a vector having compon-
ents fj(%1,%2,...,Xn,t),i=1,2,...,m. We shall assume that the f;
are continuous and satisfy standard conditions, such as having continuous
first partial derivatives so that the solution of (5.1) exists and is unique
for given initial conditions (see, for example, Brauer and Nohel 1969).
If the f; do not depend explicitly on ¢, eqn (5.1) is called autonomous
(otherwise, nonautonomous). If f(c, t) = 0 for all t, where c is some
constant vector, then it follows at once from (5.1) that if x (to) = c then
x (t) =c, all t >to.Thus solutions starting at c remain there, and c is said
to be an equilibrium or critical point (or state), Clearly, by introducing
new variables x;’ = x; —c; we can arrange for the equilibrium point to be
transferred to the origin; we shall assume that this has been done for any
equilibrium point under consideration (there may well be several for a
given system (5.1)) so that we then have f(0, t) = 0, t >to. We shall also
assume that there is no other constant solution in the neighbourhood of
the origin, so this is an isolated equilibrium point.
Example 5.1.
The intuitive idea of stability in a dynamical setting is that for
“small” perturbations from the equilibrium state at some time fo,
Stability 143
subsequent motions x(t), f> fo, should not be too ‘large’. Suppose that
Fig. 5.1 represents a ball resting in equilibrium on a sheet of metal bent
into various shapes with cross-sections as shown. If frictional forces can
be neglected then small perturbations lead to:
(iii) unstable if it is not stable; that is, there exists ane <0
such that for every 5 > O there exists an x(fo) with
llx(to) Il<5, |lx (£1) || = € for some f, > fo. If this holds
for every x (fo) in ||x (to) || <6the equilibrium is completely
unstable.
-=—
FIG. 5.1
b, which may depend on fg and x(t), such that || x(t) || <b for all
t2fo.
Regarded as a function of f in the n-dimensional state space, the
solution x(t) of (5.1) is called a trajectory or motion. In two dimensions
we can give the definitions a simple geometrical interpretation.
If the origin O is.stable, then given the outer circle C, radius e, there
exists an inner circle C,, radius 5,, such that trajectories starting within
C, never leave C. If O is asymptotically stable then there is some circle
C,, radius 62, having the same property as C, but in addition trajectories
starting inside C, tend to Oast > ~.
Some further remarks can be made at this stage:
(1) Notice that we refer to the stability of an equilibrium state of
(5.1), not the system itself, as different equilibrium points may
have different stability properties.
(2) A weakness of the definition of stability i.s.L. for practical
purposes is that only the existence of some positive 6 is required,
so 6 may be very small compared to € —in other words, only very
small disturbances from equilibrium may be allowable.
Stability 145
FIG. 5.2
where d =a, —dq, and a, /a3 = 4@/a4, with a similar expression for x(t).
If d>0,x,(t) and x2(f) tend to infinity as t > ©, so the origin is
unstable. If d<0, both x, (t) and x(t) approach constant values as
t > ©, and it is intuitively obvious that the origin is stable. This could
be verified formally by showing that the condition of the definition is
satisfied, but we shall see in Section 5.2 that linear equations like (1.6)
can be tested directly.
146 Stability
Xa = oo (5:2)
It is clear that the solution exists and is unique, and in fact by integrating
we easily obtain
—1/x, =t—-1/x?
D<(t) = eo
; (i/x})-t
xy =X el = 28) (5.3)
Xan 3
ee (5.6)
0, SP
NE) x4
where
Xa) (5.9)
mig
FIG. 5.3
The origin of these coordinates is an equilibrium point, and from (5.9)
dx, /dx ee—x,/sin xy
Exercise 5.1.
Determine the equilibrium point (other than the origin) of the system
described by
Ny HS
X1> 2X1x3
Xp ==2X5 $X1X2
FIG. 5.4
Exercise 5.2.
Show by solving the equations explicitly and applying the definition that
the origin is a stable equilibrium point for the system
Exercise 5.3.
If the oscillations of the simple pendulum in Example 5.7 are assumed
sufficiently small so that 67 and higher powers are neglected, and if
g/2 =k, obtain the equation of trajectories and sketch them in the phase
plane.
Exercise 5.4.
Verify that for the inverted pendulum problem in Exercise 1.3 the
origin is an equilibrium point.
x =Ax, (5.10)
q
exp (Af) = Zip t Zegtt ts * + Ziog tk 1) exp (Axt)
k=]
Stability 151
where the Zx¢ are constant matrices determined entirely by A. Using
properties of norms (see Section 2.7) we obtain
ial
llexp(At)I<> SSt2-1| exp(Age)
IlIlZeeIl
k=1Q=1
= > Dro Lexp[Rex)e]IhZeeIl (5.12)
= (Dewye S22
provided Re(Ax) < 0, all k, since (5.12) is a finite sum of terms, each of
which > Oast > ©. From (5.11)
A, =
0) i
(the c; are constants) so the system is clearly stable. Using the definitions
in Section 2.5 the elementary divisors of A, are found to be A-i, A-i
so 8B,=B, = 1.
If however we consider the matrix
A, =
1 i
the roots of A, are again i, i but the elementary divisors are 1, (A—i)*
so B, = 2. It is easy to verify that the solution of (5.10) in this case is
x(t) =c3 exp (it), x2(t) =cq exp (it) + c3¢ exp (it)
showing that the origin is indeed unstable.
In fact (see Exercise 2.24) the stability condition on the B; in Theorem
5.2 is equivalent to requiring that all the Jordan blocks J(\j),i=1,2,...,
in the Jordan form (2.59) of Amust have order one. This is clearly illus-
trated in the above example, where
a sa!
A =
a3 aA
det(AJ-—A) =A(A-d)
using the condition a; /a3 = 4 /a4. Hence A has a single zero root, so the
system is stable provided d (=a, —aq) is negative.
The two preceding theorems apply if A is real or complex, so the
stability determination of (5.10) can be carried out by computing the
Ax using one of the powerful standard computer programs now available.
However, if 7 is small (say less than six), or if some of the elements of
A are in parametric form, or if access to a digital computer is not possible,
then the classical results given below are useful.
Because of its practical importance the linear system stability problem
Stability . 153
has attracted attention for a considerable time, an early study being by
Maxwell in 1868 in connection with the governing of steam engines. The
original formulation of the problem was not of course in matrix terms,
the system model being (3.42). This is equivalent to working with the
characteristic polynomial of A, which we shall write in this section as
det(QQU-—A)
=X?+a,;N?~
1 ++++tan _—1d+apn
=a(A),(5.13)
and this is identical to k(A) in (3.47) when the system equation is (3.42).
The first solutions giving necessary and sufficient conditions for all the
roots of a(A) in (5.13) to have negative real parts were given by Cauchy in
1831, Sturm in 1836, and Hermite in 1854, but we give here a well-known
theorem due to Hurwitz in 1895 for the case when all the a; are real (for
a good account of the history of the problem see Fuller 1974).
Qa, a3 as aed
1 az a4 G42n-—2
0 a a3 . ; G2n—-3
PaO, eT. cal inl)< 3 wototivy
ag tg? for(i)
where a, = 0,r >n. Let H; denote the ith leading principal minor of H.
Then all the roots of a(A) have negative real parts (a(A) is aHurwitz
polynomial) if and only if H; >0,i=1,2,...,7.
We shall not give a proof, which usually involves complex variable
theory (see Gantmacher 1959, 2 pp.177, 190). However, some remarks
on an alternative way of obtaining Theorem 5.3 will be given later in
Section 5.5.
Of course a disadvantage of Theorem 5.3 is the need to evaluate
determinants of increasing order, and a convenient way of avoiding this
is due to Routh whose work was published before that of Hurwitz, in 1877.
154 Stability
1 0 0 ay, a3 0)
(- 1/a,) 1 0 1. @7
—a,/(a3 —a,a2) a," /(a3 a2) ] 0 ay a3
AH=R, (5.16)
ry ae), VAs : é 3
Vr P7923 2 ‘
{To1,
7025703;
- .# {1,42,a4,. . at
{71571257135---} {41,43rar ay,
M%-2,1 M-2,j7+1
file yet aeee eel, 20)
Fi-1,1 Ni-1,j+1
Thus each row in (5.18) is formed by taking all the 2 x 2 minors involving
the first column of the preceding two rows. This is shown by the elements
within dashed lines in (5.18). The computational effort involved is clearly
much less than for Hurwitz’s method.
As mentioned earlier, the Hurwitz and Routh theorems can be useful
for determining stability of (3.42) and (5.10) in certain cases. However,
it should be noted that a practical disadvantage of application to (5.10)
is that it is very difficult to calculate accurately the a; in (5.13). This is
‘important because small errors in the a; can lead to large errors in the
roots of a(A).
rt + 2A + 9A? +A 44.
ro] l 2 :
rj 2 I
rj (18—1)/2 (8 —0)/2
= 17/2 a4
r3j [(17/2) —8] /(17/2)
=1/17
raj [(4/17) —0] /Q/17)
=4
BPSOTSfee (5.21)
Proof. Any complex roots of a(A) will occur in conjugate pairs a + i,
the corresponding factor of a (A) being
(A —a —iB) (A —@
+ iB) = A? —20d + a? + B?.
By Theorem 5.1, a< 0, and similarly any real factor of a(A) can be
written (A + y) with y <0. Thus
For a proof see Gantmacher (1959, 2 p. 221). Notice that of course any
one of eqns (5.23)—(5.26) implies (5.21). Thus the Lienard-Chipart
criterion shows that if (5.21) holds then only 4 (n —1) or 7n of the Hj
in Theorem 5.3 need to be evaluated according as n is odd or even. In fact
it has been found recently that a reduction by half in the orders of the
determinants is also possible, and the development is interesting since it
involves companion matrices and is linked with the ideas of Section 4.3.
We have no room here to give details and refer the reader to Barnett
(1971, Chapter 2). We have mentioned earlier that for numerical
calculations the Routh array is preferable to the Hurwitz determinants.
However, if some of the a; depend upon parameters and the system is
asymptotically stable it is often required to determine for what values of
_ the parameters a (A) first becomes non-Hurwitzian. It can be shown
(Lehnigk 1966, p. 199) that these values are obtained by solving the
‘critical’ equations.
@n=0, Hny-1=0. (5.27)
Although encountered far less often in practice we also give a result for
the case when the a; in (5.13) are complex.
are positive.
For a proof and other results on complex polynomials see Marden
(1966, p. 179).
158 Stability
5.2.2. Discrete-time
We now turn our attention to discrete-time linear systems in the usual
form
x(k +1)=A,x(k). (5.30)
0.12 0.95 0 0
Ay 0 0 0 1
0.14 0 0 0.95
It is easy to obtain
det ( —A,)=A(A—0.95) (A? —0.95A —0.12)
=A(A- 0.95) (A- 1.063) (A + 0.113)
so the system is unstable since A, has a root greater than unity.
Just as in Theorem 5.2, if any of the characteristic roots of Ay have
unit modulus then the corresponding elementary divisors must be linear
for (5.30) to be stable.
Stability 159
In fact we can establish a direct relationship between Theorems 5.1
and 5.8 by taking
Mia (Ay 1) (A; +1) * (5.31)
Qo Qy : C ¢ O—-1 ON
On On —1 o as Oy Xo
dy dz2 : E ¥ dan
C31 C32 = i C3>n-1
(Sis Be Sut
Kili ihe HO Ci-1,n—i+3
ps otiili acer eco
paagee
aeei
y : Cj Cig +1 :
160 Stability
and
nes Cit da cled i= 2,3,...,% #9 (5.36)
“ij Gi-i,1 Ge a,p44}?
Then the numbers of roots of a, (A) inside and outside the unit disc are
Nand n-—VN,where N is the number of negative products Py defined in
(5.39). For further details and other results, including a determinantal
criterion corresponding to that of Hurwitz in Theorem 5.3, the book by
Jury (1964, Chapter 3) should be consulted.
Cj 2 4 7) 3
dij 3 2) 4 2
C2;
dj =22
=> as)
23 ma)|
ree
C3; cae: ai
d3; 459 Sal
Cay —57800
= da
Thus in (5.39), P; > 0, P, <0, P3 <0 so (5.40) has two roots inside
the unit disc and one outside, and is therefore the characteristic
equation of an unstable discrete-time system by Theorem 5.8.
When the a; in (5.33) are complex the only change required is that
the second row of each pair in (5.35) is the complex conjugate of the
| preceding row in reverse order.
x (t)=A(t)x(t) (5.41)
where A oois a constant matrix, then the following result can be establish-
ed.
THEOREM5.10. If the origin as an asymptotically stable equilibrium
point for the system
iael
Sd
162 Stability
Exercise 5.5.
Analyse the problem in Exercise 5.2 using Theorems 5.1 and 5.2.
Exercise 5.6.
Determine whether the following polynomials are Hurwitzian:
(aq) 8 +1727 +2041 (b) 04 +23 +40? +4043
(c) 044643427 +A43 @ 405 +5074+2602 430K IX + 2.
Exercise 5.7.
Determine the solution & (f) of the linear system
Exercise 5.8.
Determine the location of the roots of
pt +2? +p? +3ut+2
Exercise 5.9.
Show that necessary and sufficient conditions for the real polynomial
a(A) to be Hurwitzian are (a) when n = 2: a; > 0,a, >0(b) when
n=3:a, >0,a3 >0,a,a, —a3 >0.
Stability 163
Exercise 5.10.
Show that necessary and sufficient conditions for the real polynomial
a;(u) to be convergent when n = 2 are a, —A <0,a,(1)>0,
a; ew:
Exercise 5.11.
Determine for what range of values of k the polynomial
(3 —kK)A?+ 2d? + (5-2k)A +2
is Hurwitzian.
Exercise 5.12.
Test the polynomial
AP + (1 +i)? +(2-3IA +7,
Exercise 5.13.
The characteristic equation
d? + 2kwAr+ (w? —2kwMi) = 0,
Exercise 5.14.
Determine for what range of values of k the real system
) ] @)
=H =! =)
i u
is added, find a linear feedback control which makes all the characteristic
roots of the closed loop system equal to — 1.
164 Stability
Exercise 5.15.
Verify that the characteristic roots of
-4 oC.
A(t)=
—e%! 0
are both constant and negative, but that the solution of (5.41) diverges as
[ieree?ok
Exercise 5.16.
Define a Routh array (s;;) without divisions by soj = oj, S17 ="17
Si-2, 1 Si-2,j+1
Sa : i=2;53.00
Si-1,1, Si-1,j+1
where the 7j; are given by (5.19) and (5.20). Show that this corresponds
to taking 827 =/oj, S37="ualsj, Sig= Kitz, 1> 2 where
k; = kj_ki_ 271-1, 1,12 4. Hence deduce that the Hurwitz determin-
ants are given by H, =54,,H» =821,H3 =53,, Hj = 5j1/8;, i > 4 where
Ne | Sy, (i even)
i**i-1 Si-3,1Si-5,1 S11(i odd).
60=4 (5.49)
The mapping of the imaginary axis s = iw, -° < w < is given by
Ut+iV=-
iw t+]
so
U=(1+w?)"?, V=-w(lt+w?)!
(U-2) as = 2.
This shows that the mapping of the imaginary axis in the s-plane is a
circle in the g(s)-plane, centre 3 + i0, radius 2 (see Fig. 5.5). As w
goes from zero to © the semicircle from A to B is traversed. Since
it follows that as w goes from —© to zero the path is simply the complex
conjugate of BA, shown by the dashed curve in Fig. 5.5.
166 Stability
s-plane g(s)-plane
FIG. 5.5
FIG. 5.6
Stability 167
Any point on the curved part of the contour can be written in polar
form as s =r exp(i@) and substitution into (5.47) gives
g(s)=B(re!®)/k(re!®). (5.50)
If we assume, as usually happens in practice, that g(s) is proper, then
58 < 8k and in (5.50) | g(s) |> 0 asr > ©. Thus the semicircle ofinfinite
radius is mapped into the origin of the g(s)-plane. We need therefore
only consider the imaginary axis in Fig. 5.6, and the Nyquist locus
(or diagram) of g(s) is defined as the curve
FIG. 5.7
168 Stability
The Nyquist locus is now the path traced out by g(s) as s traverses the
indented imaginary axis.
Example 5.14. Consider
1
a(s) = s(s + 1)
(a) (b)
On be, s = iw so that
g(s)= eer
giving
U=-(1+w?)!, 9 V=—-{w(1
+o)?
whichisthepathb’c’in Fig.5.8(b),Onab,s =pexp(id)withp> 0, so
Rim
cues p>0 pel®(1+pel?)
lim 1
pe pel?
Stability 169
Thus as p > 0, g(s) has modulus 1/p, argument —0 with O<@<7n/2. It
follows that a’b’ is the mapping of ab. The reflection in the U axis,
corresponding to the path eda, is shown in dashed lines. The point
—1 +10 is not encircled so the system is asymptotically stable. Again it is
easy to verify that
ee a
Eels)= ststl1-
and the characteristic polynomial is asymptotically stable (see Exercise
5.9(a)).
Clearly it is sufficient to sketch only the salient features of the Nyquist
locus, a time-consuming detailed diagram usually not being necessary.
A result more general than Theorem 5.11 for the case when the open-
loop system is not asymptotically stable can also be proved. The argument
is based on Cauchy’s residue theorem and can be found in most textbooks
on classical linear control theory.
THEOREM 5.12. The closed loop system with transfer function (5.46)
is asymptotically stable if and only if
N=-PS<0
FIG. 5.9
FIG. 5.10
If a circle with unit radius and centre the origin cuts the locus at the
point P, then the whole Nyquist locus for k = 1 can be rotated clockwise
through 70° before the point —1 + i0 is reached and the system is no
longer asymptotically stable.
The quantities 1/(4) = 4 and 70° are called the gain and phase margins,
and thus provide a measure of how far the system is from instability (in
this case when k = 1). In practice a gain margin of at least two and a
phase margin of at least 30° are desirable as ‘safety factors’.
The Bode plot (mentioned in Section 1.2.1) can also be used to det-
ermine gain and phase margins, since it presents essentially the same
information as the Nyquist diagram in a different form.
In general, if g(s) is replaced by kg(s) in (5.46) the point —1 + i0 in
the g(s)-plane in Theorems 5.11 and 5.12 is replaced by —1/k + i0.
It is worth mentioning that in Fig. 5.9 the ideas of gain and phase
margins are not very meaningful. For example, the system becomes
unstable if the gain is increased so that the point A is to the left of
—1, or if the gain is decreased so that B is to the right of —1. Such a
system is called conditionally stable.
172 Stability
and the Nyquist contours of t,(s) are called the characteristic loci, It can
be shown that the necessary and sufficient condition for closed loop
asymptotic stability becomes
m
>»nas Py
k=1
where nx is the number of clockwise encirclements of —1 + i0 by the
Nyquist locus of t;(s).
When H(s) =Jm in (4.40) the closed loop transfer function matrix is
Exercise 5.17.
Verify by substituting (5.47) into (4.41) that the characteristic equation
of the closed loop system is B(s) + k(s).
Exercise 5.18.
Determine the number of clockwise encirclements by the contour shown
below of the points (a) the origin (b) —1 + iO.
FIG. 5.11
Exercise 5.19.
Sketch the Nyquist diagram for
g(s)=
5. 1)
and hence determine the number of closed loop poles in Ty’.
174 Stability
Exercise 5.20.
Sketch the Nyquist locus for the open loop transfer function
Qs ae
& s*(s - 1) ’
k
8) =(4 10)
Hence calculate the gain and phase margins when k = 800/3.
Verify that your result for the gain margin is correct by applying the
Routh criterion to the closed loop characteristic polynomial.
subject to x(to) = Xo, but modifications needed to deal with (5.1) are
straightforward. The aim is to determine the stability nature of the equi-
librium point at the origin of (5.53) without obtaining the solution x(t).
This of course has been done algebraically for linear time invariant
systems in Section 5.2. The essential idea is to generalize the concept of
energy V for a conservative system in mechanics, where a well-known
result states that an equilibrium point is stable if the energy is aminimum.
Thus V is a positive function which has V negative in the neighbourhood
of a stable equilibrium point. More generally, we define a Liapunov func-
tion V(x) as follows:
(i) V(x) and all its partial derivatives 0V/dx;are continuous.
(ii) V(x) is positive definite, i.e. V(0) =0 and V(x) > 0 forx #0
in some neighbourhood || x || <k of the origin.
(iii) the derivative of V with respect to (5.48), namely
Proof of Theorem 5.13. This relies on the fact that because of the sign
property of V there exists a continuous scalar function $(7) of r which
vanishes at r = 0 and increases strictly monotonically? in0 <r<k such
that
V[x(t1)] <VEeCo)],t1 2 to
< $(€), by (5.56). (5.57)
If there exists some f, >to such that || x(t, ) || >€ then by (5.55)
Vix(t)) >oil x(t) Il
> > (€)
which contradicts (5.57). Hence for all t; > fo we have || x(t, ) || <<e, so
the origin is stable since 6 = € in the definition in Section 5.1.
The proof of Theorem 5.14 is similar (see Hahn 1963, p. 15). It is
worth noting that if the conditions on V in Theorem 5.14 hold every-
where in state space it does not necessarily follow that the origin is
asymptotically stable in the large. For this to be the case V(x) must
have the additional property that it is radially unbounded, which means
that V(x) > © for all x such that || x || > °°. For instance, V=xj{ + x3
is radially unbounded, but V=xj /(1 + x7) + x3 is not since, for
example, V> 1 asx; >°,x 0. A similar line of reasoning shows
that if Q is the set of points outside a bounded region containing the
origin, and if throughout 2, V >0, V <0, arid Vis radially unbounded
then the origin is Lagrange stable (Willems 1970, p.34).
It may be helpful to consider a geometrical interpretation of Theorem
5.14 when n = 2.
FIG. 5.12
FIG. 5.13
If first the spring is assumed to obey Hooke’s law then the equation of
motion is
Ztkz=0, (5.58)
where k is the spring constant. Taking x; =z, x2 =Z, (5.58) becomes
DeeSa xs = ho: (5.59)
Since the system is conservative the total energy
=4 kx? +4 x2 (5.60)
is a Liapunov function and it is very easy to see from (5.59) that
E = kx, (x2) +x2(—kx) =0,
so by Theorem 5.13 the origin is stable (of course this is trivial since
(5.58) represents simple harmonic motion). Suppose now that the force
exerted by the spring, instead of being linear is some function k(x, )
satisfying k(0) = 0, k(x, ) #0 if x, #0, so that the second equation in
(5.59) becomes x, = —k. The total energy is now
E=4x2+
fc k(a)do (5.61)
and
E=x,(-k) +kx, =0,
so again by Theorem 5.13 the origin is stable for any nonlinear spring
satisfying the above conditions.
Example 5.18. Consider now the system of the previous example but
with a damping force dz added, so that the equation of motion is
Z+ditkz=0 (5.62)
178 Stability
and the total energy is still Fin (5.60), so that using (5.63)
=—x3
From (5.63)
V= 14x43
(2) + 2[x2(2) +%1©24y ee
+.6X9)(S2X4)ee)
=—4x? —4x3,
ty =exepPxge-*k =i (5.65)
Stability 79
so if F is defined as in (5.61),
E=x,(-k-dx,)+kx,
=-—x3d<0.
u=—z+e1—z7)z
RE NXo
Xe =H —(x7 —1)ra. (5.67)
V= par ar + 2x2X2
= 2ex3 (1 —x?).
Thus V <0ifx,? <1, and then by Theorem 5.15 the origin is asympto-
tically stable. It follows that all trajectories starting inside the region
Tl:x? +x3 <1 converge to the origin as t > ©, and I’ is therefore called
a region of asymptotic stability. The reader may be tempted to think
that the infinite strip x} < 1 is a region of asymptotic stability. This is
not in fact true, since a trajectory starting outside [ can move outside
this strip whilst continuing in the direction of decreasing V circles, and
hence lead to divergence (see Fig. 5.14).
180 Stability
FIG. 5.14
=-x3—€(} 2° —2)
aes -e(3 x} —X)).
HenceusingV=x7+x3,
V=2x, (—x3
—3ex? +ex,) +2x3x,
=2ex}(1-3 x7)
<0 if x7<3,
so the region of asymptotic stability obtained by this different set of
state variables is x} + x3? <3, larger than before.
Stability 181
In general if the origin is an asymptotically stable equilibrium point
then the total set of initial points from which trajectories converge to
the origin as t > © is called the domain of attraction. Knowledge of this
domain is of great value in practical problems since it enables permissible
deviations from equilibrium to be determined. However, Example 5.20
illustrates the fact that since a particular Liapunov function gives only
sufficient conditions for stability, the region of asymptotic stability
obtained can be expected to be only part of the domain of attraction.
Different Liapunov functions or different sets of state variables may
well yield different stability regions. We shall discuss methods of con-
structing Liapunov functions for nonlinear systems in Section 5.6, but
the general problem of finding ‘optimum’ Liapunov functions, which
give best possible estimates for the domain of attraction, is a difficult one.
It may be a waste of effort trying to determine the stability properties
of an equilibrium point, since the point may be unstable. The following
theorem is then useful:
Exercise 5.22.
Write the system
2+2z+27=0
in state space form. Let
V =axt+ bx? + 0x x, + dx}
and choose the constants a, b, c, d such that
7 = —x4 —x2,
Exercise 5.23.
Using the function
Exercise 5,24,
Investigate the stability of the origin for the system
Exercise 5.25.
For the equations in Exercise 5.1, investigate stability of the origin
using V= x3 —x?.
If in Exercise 5.1 the transformed system variables are y, and y2,
use
V=yy+2y2—2kn
(1+ 41) —-Ln
(1+2y2)
to investigate the stability of the second equilibrium point.
Stability 183
Exercise 5.26.
Use the function V =x? +x} to show that the origin of the system
ero eet 2x2, X_ =-2x1 —X2
Exercise 5.27.
Find the domain of attraction for the origin of the system described by
the scalar equation
2=% 2(z—2).
Exercise 5.28.
The equations of motion of a gyroscope without external forces are
Aw, +(C-B)w,w; =0
Bo, +(A-C) w30, =0
Ca 3 + (B-A) w, Ww,=0
where A, B, C are the principal moments of inertia and w,, w2, Ws are
the angular rates about the principal axes. Put w, = Wo +X1, W2 =X2,
W3 =X3 (where Wo is a constant) and show that the origin of coordinates
is then stable if A < B<<C by using the function
Exercise 5.29.
Show that if the origin is an asymptotically stable equilibrium point for
(5.53) with Liapunov function V (x), then it also is for
Apply this result to the system in Exercise 5.23 by choosing S=0 and
a suitable matrix Q(x) to show that the second equation can be replaced
by
F2=—xX, —X2+ (&, + 2x2)[h@1,%3) +25
X=Ax. (5.70)
ViextPx (S.71)
where P is a real symmetric matrix. The time derivative of Vwith respect
to (5.70) is
where
A'P+PA=-Q, (5.72)
and it is easy to see that Q is also symmetric. If P and Q are both positive
definite then by Theorem 5.14 the (origin of) system (5.70) is asymptot-
ically stable. If Q is positive definite and P is negative definite or indefinite
then in both cases V can take negative values in the neighbourhood of the
origin so by Theorem 5.16, (5.70) is unstable. We have therefore proved:
V=—xTQx
+27 Px. (5.78)
Because of the nature of g, the second term in (5.78) has degree three at
least, and so for x sufficiently close to the origin, V<<0. Hence by Theorem
5.14 the origin of (5.76) is asymptotically stable.
Conversely, if (5.77) is unstable, V remains negative definite but P is
indefinite so V can take negative values and therefore satisfies the condi-
tions of Theorem 5.16 for instability.
In fact the condition on g can be relaxed to || g(x) || <A || x || for
sufficiently small \ > 0 (Lehnigk 1966, p. 55). Notice that if (5.77) is
stable but not asymptotically stable, Theorem 5.19 provides no informa-
tion about the stability of the origin of (5.76), and other methods must
be used.
Furthermore, it is clear that linearization cannot provide any informa-
tion about regions of asymptotic stability for the nonlinear system (5.53),
since if the first approximation (5.77) is asymptotically stable then it is
so in the large. Thus the extent of asymptotic stability for (5.53) is
determined by the nonlinear terms in (5.76).
188 Stability
or
Ngee
XoswXe, =—bx ypsexes Oi, xo (5.79)
The linear part of (5.79) is asymptotically stable if and only ifa >0,
b > 0 (see Exercise 5.9), so if g is any function of x, and x2 satisfying
the conditions of Theorem 5.19, the origin of (5.79) is also asymptotically
stable.
Exercise 5.30.
Using the matrix A in Exercise 3.1(a) and taking
Pi P2
P2 P3
solve the equation (5.72) with Q =/. Hence determine the stability
nature of A.
Exercise 5.31.
Integrate both sides of the matrix differential equation in Exercise 3.12
with respect to t from t = 0 to t = %°, Hence deduce that if A is a stability
matrix the solution of (5.72) can be written as
P=(e exp
(A'L)Q
exp
(Af)dt.
Exercise 5.32.
Convert the second order equation
Za+a,Z + azZ =0
into state space form. Using V in (5.71) with V =—x3obtain the necessary
and sufficient conditions a; >0,a, > 0 for asymptotic stability (see .
Exercise 5.9).
Stability 189
Exercise 5.33.
By using the quadratic Liapunov function V(x) which has derivative
—2(x} + x3), determine the stability of the system
Exercise 5.34.
Write (5.72) in the form
(PA +3Q)+(PA+4Q)' =0
Exercise 5.35.
Verify that the matrix equation (5.74) has solution
Exercise 5.36.
Replace A by A + o/ in Theorem 5.17, where o is a real number. Hence
deduce the relative stability condition that the real parts of the roots of
A are less than — o if and only if the solution P of
A™P+PA+20P=-Q
is positive definite.
Exercise 5.37.
Use Theorem 5.17 to show that the matrix A = Po(So —Qo) is a stability
matrix, where Pp and Qo are arbitrary r.s.p.d. matrices, and So is an
arbitrary real skew symmetric matrix.
190 Stability
Exercise 5.38.
If for a given real stability matrix A the solution of (5.72) is P, show that
A +(S; —Q,)P is also a stability matrix, S, and Q; being arbitrary real
skew and positive semidefinite n x n matrices respectively. (Compare with
Exercise’ 5.29):
Exercise 5.39.
If n = 2, show that eqn (5.30) is asymptotically stable if and only if
|det A, |1<1 and 1 + det A, >| trA,|. (Hint: use the result in Exercise
5.10).
What is the corresponding result for eqn. (5.70)?
Exercise 5.40.
The equations describing a time varying LCR series electric circuit are
X, =X2/L(t), X2. =-x1/C(H)-—x2RW/LO,
where x, is the charge on the capacitor, x2 is the flux in the inductor
and L(t), C(t), R(t) >0,¢t 20.
Use the Liapunov function V = x" P(t)x, where
R+2L/RC 1
P(t) =
- 1 2/R
to show that V = —x™Q(t)x, where
AIC. 0 ,
Q(t) = ar
0 2/L
and hence deduce that sufficient conditions for the origin to be asympto-
tically stable are
1+R(L/R? -$C)+CL/RC-L/R>O,
1+RL/R?>0
(it can be assumed that Theorem 5.14 still holds even though the system
is non-autonomous).
Notice that it would be difficult to obtain ‘sharp’ (i.e. almost necessary)
conditions without considering the nature of the functions L(t), C(t),
R(t) in detail.
Stability 191
Exercise 5.41.
Investigate the stability nature of the equilibrium points at the origin for
the systems
(a) %, =7x,+2sinx, —x}
a= exp (x4) —3X2—1 +:5x7
(b) X1 =(3/4)sinx, —7x2 (1-x2) 5 +x3
a= 2/3) x1 —3x2 Cos'xg 11x37:
Exercise 5.42.
In the inverted pendulum problem of Exercise 1.3 assume that motions
about equilibrium are so small that second and higher degree terms in
the Taylor series expansion of f(x, xX)can be neglected, and hence obtain
the linearized form of the state equations. By finding the characteristic
equation, use Theorem 5.19 to determine the stability nature of the
origin (see Example 1.4 and Exercise 5.4).
Exercise 5.43.
Solve the scalar equation
z=z-etz3
by using the substitution z = e'w. Hence verify that z(t) > 0 as t > ©,
although the linear part is unstable.
This illustrates that Theorem 5.19 does not carry over to linear time
varying systems.
We have seen in the previous section how a quadratic form can always be
used as a potential Liapunov function for a constant linear system. Before
seeking Liapunov functions for the nonlinear system (5.53) we must be
certain that we are not wasting our time, and this is ensured by a number
of results which are the converse of the stability theorems in Section 5.4.
We quote one of these as an example, and refer the reader to Hahn (1963,
Chapter 4) for further details.
oV OV
PestsEiglace
td oy aV
ax,1”
=(VV)'f (5.80
where
Qa ky eee
Se Bae
VV = . (5.82)
peegna
xn: z wn Onna
the a; being functions of the xs and chosen so as to satisfy the conditions:
(1) Vin (5.80) is negative definite
(2) VV in (5.82) is indeed the gradient of a scalar function.
From vector theory this requires that the n-dimensional curl of VV be
identically zero, i.e.
aG;_ 9G;
wal
pelea de, Delletel
nty 5.83
(5.83)
G#/),
where G; is the ith component in (5.82).
After the conditions (5.83) have been satisfied, acheck is then needed
Stability 193
to ensure that Vis still negative definite —if not, the procedure will have
to be repeated. Finally, V is obtained from the line integral
[o'GiGe1,0,
0,.oe,O)dx,
1 yeG2(X1,X2,0,.
o-, O)dx,
feoe t+ fin Gn(%15,X2;
besXn dxn: (5.84)
Of course the n? functions aj; are not determined uniquely by the above
conditions so to that extent the method is one of ‘trial and error’. Note
also that failure to find a suitable function using the variable gradient
method does not imply anything about the stability nature of the equi-
librium point.
0 =—3
0 a=
which has characteristic roots 0 and —2, and so is only stable (Theorem
5.2). Hence the linearization Theorem 5.19 does not apply.
From (5.80) and (5.82)
Q12=0, G1 =0,
a Oy M
= § aneXt.
4
194 Stability
This gives
V=—01x§ —2ay,x?
5 On X} G,
Vics =
Q22X2 G,
V=\,Xi3QpQx7
4 5 dx,+ifXo
92X72
dxX2
= Q(x} + 9x3)/18
The aim is again to try and find a Liapunov function by starting with a
negative definite function ¢(x) as its derivative, so we now wish to solve
for V(x,,X2,...,Xn) the partial differential equation
OV
et het da ot ee ee (5.86)
Oxy 0x2 OXp
subject to the boundary condition V(0) = 0. Equation (5.86) can be
written
dV/dt =
which on integration with respect to ¢ gives
VEe(T)]
~Vero)
=f° ole(Oldr, (5.87
Stability 195
where Xo =X(to ). If the origin of (5.53) is asymptotically stable, and
Xo lies within its domain of attraction then letting T > © in (5.87)
produces
Vito}
=—ikedt
which is positive, so we can expect the solution of (5.86) to be positive
definite. As Xo approaches the boundary of the domain of attraction then
in (5.87) V(xq) tends to infinity since Vix(T)} > occas T > °°, To avoid
this difficulty the procedure can be altered by defining a function
Thedifferential
equation
(5.86)
thenbecomes
Lae
axa” axe*7 een
Onee
a =e Wye
Ke 5:99
CE
The boundary of the domain of attraction is now given by W(x) = 1.
Exercise 5.44.
Consider the equilibrium point at the origin for the system
Exercise 5.45.
In Exercise 5.22 deduce the stated expression for V using the variable
gradient method.
Exercise 5.46.
The equations describing a homogeneous atomic reactor with constant
power extraction can be written
X, =—ax,/7,
ty ={exp(r1)- Ile
where P(t) = exp x, (t) is the instantaneous reactor power, x, (f) is the
temperature, a > 0 is the temperature coefficient, e> 0 is the heat
capacity and 7 > O the average life of a neutron. Use the variable gradient
method with a2 = 1, 2 = 21 = 0 in (5.82) to determine the stability
nature of the equilibrium point at the origin. (Hint: set V = 0).
Exercise 5.47.
Consider the first order system
x1 mre mds
Use Zubov’s method with ¢ =— 2x? to determine the stability nature of
the origin.
Apart from Section 5.3, where we developed the Nyquist criterion for
closed loop linear systems, our discussions in this chapter have so far
not directly involved the control terms in the system equations. We now
consider some stability problems associated explicitly with the control
variables.
x =f(x,u,t), f(0,0,t)=0
with output y = g(x, u, f) is said to be bounded input —bounded output
Stability 197
where , is any positive constant, then there exists anumber 2, > 0 such
that || y(t) || <2, for t > to, regardless of initial state x(t). The problem
of studying b.i.b.o. stability for nonlinear systems is a difficult one, but
we can give some results for the usual linear system
x =Ax + Bu (5.91)
y=Cx. (5.92)
Using (3.22) and the properties of norms (see Section 2.7) we have
Iy@ ISIC Il tI
THEOREM 5.22. If the system (5.91) and (5.92) is c.c. and c.o. and
b.i.b.o. stable, then x = Ax is asymptotically stable.
For a proof see the book by Willems (1970, p. 53). Similar results for
discrete-time systems also hold (see Exercise 5.51). However, for linear
time varying systems Theorem 5.21 is not true (see Exercise 5.52) unless
for all t the norms of B(t) and C(t) are bounded and the norm of the
198 Stability
with |u |<1 is not c.c. by Theorem 5.23, but if the second equation is
Consider again the linear system (5.91). If the open loop system is
unstable (i.e. by Theorem 5.1 one or more of the characteristic roots of
Stability 199
A has a positive real part) then an essential practical objective would be
to apply control so as to stabilize the system —that is, make the closed
loop system asymptotically stable. If (5.91) is c.c. then we saw in Section
4.4 (Theorem 4.17) that stabilization can always be achieved by linear
feedback u = Kx, since there are an infinity of matrices K which will
make A + BK a stability matrix. If the pair [A, B] is not c.c. then we can
define the weaker property that [A, B] is stabilizable if and only if there
exists a constant matrix K such that A + BK is asymptotically stable. We
can see at once from the canonical form of (5.91) displayed in Theorem
4.9 that the system is stabilizable if and only if A3 in (4.21) is a stability
matrix. In this case the feedback is u = K,x), where K, can be chosen
to make A, + 8B,K, asymptotically stable since the pair [A,, 8, ] isc.c.
By duality (see Theorem 4.11) we define the pair [A, C], where C is the
output matrix in (5.92), to be detectable if and only if [47, C™] is
stabilizable.
Methods for constructing feedback matrices K were discussed in
Section 4.4, where the disadvantages of the method of prespecifying
closed loop poles were indicated. Two simple illustrations of the applica-
tion of stabilizing linear feedback have been provided by the rabbit—fox
problem (Exercise 4.33), in which a disease lethal to rabbits keeps the
animal population finite; and by the buffalo model (Exercise 4.53) in
which slaughtering animals for food fulfils the same function. Another
method of constructing matrices K will be developed in Chapter 6 using
the ideas of optimal control. It is interesting to consider here a simple
application of the Liapunov metheds of Section 5.5. First notice that if
(5.70) is asymptotically stable with Liapunov function V = x" Px, where
P satisfies (5.72), then
V/V=—x"!Qx/x'Px
<-o (5.96)
wheregis the minimumvalueof the ratio x Qx/x! Px(in fact this is
equal to the smallest characteristic root of QP~'). Integrating (5.96)
with respect to ¢ gives
V[x(t)] <exp ~ ot)V[x(0)]. (5.97)
Since V[x(t)] > 0 as t >, (5.97) can be regarded as a measure of the
way in which trajectories approach the origin, so the larger o the ‘faster’
does x(t) > 0. Suppose now we apply the control
u=(S—Q,)B™Px (5.98)
200 Stability
X= [A+B(S—Q,)B'P]x (5.99)
and it is easy to verify that if V=x! Px then the derivative with respect
to (5.99) is
V=—x™Qx—2xTPBO,B™Px
<-x'Qx
since PBQ ,B' P = (PB)Q ,(PB)' is positive definite. Hence by the argu-
ment just developed it follows that (5.99) is ‘more stable’ than the open
loop system (5.70), in the sense that trajectories will approach the
origin more quickly. Of course (5.98) is of rather limited practical value
because it requires asymptotic stability of the open loop system, but
nevertheless the power of Liapunov theory is apparent by the ease with
which the asymptotic stability of (5.99) can be established. This would
be impossible using the classical methods of Section 5.2, requiring
calculation of the characteristic equation of (5.99). Furthermore, the
Liapunov approach often enables extensions to nonlinear problems to
be made (see for example Exercise 5.53).
x = Ax —bu (5.100)
subject to nonlinear feedback
u=f(y) (5.101)
where
y=cx. (5.102)
In (5.101) f(v) is a continuous function of the scalar output y, and A, b,
and c in (5.100) and (5.102) are such that the triple {A, , c} is a minimal
realization of a scalar transfer function g(s). It is assumed that f(0) = 0
so the origin x = 0 is an equilibrium point for (5.100) and (5.101).
The stability theorems to be quoted in this section involve the notion
of a positive real function r(s), which we define to be a rational function
Stability 201
p(s)/q(s) with real coefficients such that p(s) and q(s) are relatively prime
and
(a) r(s) has no poles in Re(s) > 0
(b) any purely imaginary poles of r(s) are simple and have real
positive residues
(c) Re{r(iw)}>
2 0 forallreal w>
Wecannowstate
arecent
result
athegives
asifficient
stability
condition.
THEOREM 5.24. (Popov). The origin of the system described by
(5.100), (5.101) and (5.102) is asymptotically stable in the large if
(i)
i
0< fy)
. <ky,ally£0, (5.103)
5.103
xTPx+0 Ik f(o)do
0
where P is the solution of the Liapunov matrix equation (5.72) (note
that positive-realness of (5.104) ensures that A has no characteristic
roots with positive real parts). Another approach to the proof uses
functional analysis (see Willems 1970, p. 152).
Notice that (5.103) means that provided condition (i) of Theorem
5.24 is satisfied, the origin is asymptotically stable in the large for any
continuous function f(y) lying within the sector bounded by the line
z = k,y and the y axis, as shown in Fig. 5.15. In this case the system is
said to be absolutely stable in the sector [0, k,].
An interesting and useful aspect of the Popov criterion is that it has a
simple graphical interpretation. Positive realness of (5.104) requires that
g(s) must have no poles in Re(s) > 0, and that any imaginary poles of
g(s) are simple and the corresponding residues of (5.104) are real and
- positive. In particular, if A is a stability matrix (i.e. the open loop system
is asymptotically stable) these conditions will certainly be satisfied.
allen
—
202 Stability
FIG. 5.15
which differs from the Nyquist locus (5.51) only in that the imaginary
part has an extra factor w. Substitution of (5.106) into (5.105) produces
av, Ol < 1/k,.
FIG. 5.16
Stability 203
The value of a is given by the slope of the straight line a V; —U, = 1/k;
(the Popov line). The largest value of k; would be when —1/k, + i0 is
at the point A in Fig. 5.16. If k, is increased further so that the point B is
reached (5.105) cannot be satisfied (although this does not mean the
origin is unstable, as Theorem 5.24 is only sufficient).
When f is a linear function then we simply have linear feedback, so
u=Ky and the Nyquist criterion of Theorem 5.11 (in the modified form
of p. 171) applies. The necessary and sufficient condition for absolute
stability in the sector [0, k,] (ie.0<x <k;,) is that the Nyquist locus
of g(s) should not intersect the real axis to the left of —1/k, + iO (includ-
ing the point itself), for if it did there would be some point —1/kp + i0
with kg < k, enclosed by the locus. If the system with u given by (5.101)
is to be absolutely stable, it clearly must be stable for the particular case
when fis linear, so the condition on the Nyquist locus is a necessary one
for absolute stability of the nonlinear system in the sector [0, k, ].
FIG. 5.17
204 Stability
u=h(x)cx (5.107)
is positive real.
We also omit this proof, which can be developed either by Liapunov
theory (Willems 1970, p. 161) or using functional analysis. Instead we
shall again develop a graphical interpretation using the frequency transfer
function.
The positive realness condition (c) on p. 201 applied to (5.109) requires
that
ce[tt]
pete)
Neenae |
e119
for w> 0, and this time the standard Nyquist locus (5.51) is used.
Substituting g(iw) = U + iV into (5.110) reduces it to
This circle @ has the points —1/k3 + i0,— 1/k, +i0 at opposite ends
of a diameter, as shown in Fig. 5.18.
Stability 205
FIG. 5.18
28(s)
A +) k(s) (5.112)
must be of Hurwitz type. The zeros of (5.112) are the poles of
B(s) é g(s)
and the condition that they should all be in Re(s) < 0 is (by Theorem
5.12) that the complete Nyquist locus of g(s) for —°° <w < should
encircle the point —4(1/k, + 1/k3) + i0 in an anticlockwise direction
P times, where P is the number of poles of g(s) in Re(s) > 0. This point
is just the centre of the circle @ , so when kz and k3 have the same sign
206 Stability
Exercise 5.48.
Consider the mass-spring system in Example 5.17, which is stable but
not asymptotically stable. If a vertical control force is applied to the
mass, determine whether the system is b.i.b.o. stable.
Exercise 5.49,
Consider again the inverted pendulum of Example 1.4 and Exercise 1.3,
and use the linearized model developed in Exercise 5.42 in the following:
It is required to keep the rod as nearly vertical as possible by means of
the horizontal control force u(t) applied to the platform. Determine the
control terms in the linearized system equations (see Fig. 1.5). Show that
stabilization cannot be achieved by linear feedback u = k6, but that if
k >(m + M)g then the rod can be given an oscillatory motion about the
vertical position.
What would be the implications of Theorem 4.17 for this problem?
Exercise 5.50.
Show that the system described by
0 A! 0 0
yal 0 OS Xehe sO ae pie
Stability 207
is stable i.s.L. and b.i.b.o. stable, but not asymptotically stable. (It is
easy to verify that the system is not c.c., so Theorem 5.22 does not apply).
Exercise 5.51.
Prove the analogue of Theorem 5.21 for the discrete-time system
Exercise5.52.
Considerthe systemdescribedby the scalarequation
¥1()=— x1 (H(t +3) +uf), x1(0) =Xo.
Show that if u(t) = 0, ¢ > O then the origin is asymptotically stable, but
that if w(t) is the unit step function (defined in (1.16)) then im x(t) =~.
foo
"Exercise 5.53.
If the origin of the system described by x =f(x) is stable, with Liapunov
function V(x), show that the system x = f(x) + u is made asymptotically
stable by taking
Exercise 5.54.
Consider the equilibrium point at the origin for the gyroscope system of
Exercise 5.28 with external forces, so that the equations of motion are
Exercise 5.55.
If in Theorem 5.24 g(s) = 1/(s* + as* + bs) where a > 0, b> 0, show that
the sector of absolute stability is [0, ab].
Exercise 5.56.
By considering the Nyquist locus of r(s) = p(s)/q(s) deduce from Theorem
5.11 that if7(s) is positive real and has no purely imaginary poles then
p(s) + q(s) is aHurwitz polynomial.
Exercise 5.57.
Sketch the Nyquist locus for g(s) = 1/(s? + 2s). By finding the largest
circle centred on —1 + iO in the g(s) plane which satisfies the conditions
of Theorem 5.25, show that the resulting bounds (5.108) are
2(2 —/3) < h(x) < 22 + V3).
6 Optimal control
= Mdt (6.3)
where f; is the first instant of time at which the desired state is reached.
210 Optimal control
cT+4T* =atbT?.
This equation may not have any real positive solution; in other words,
this minimum-time problem may have no solution for certain initial
conditions.
eN(t,)
Me(t1), (6.4)
where e(t) = x(t) —r(t) and M is a real symmetric positive definite
(1.s.p.d.) n x n matrix. A special case is when M is the unit matrix and
(6.4) reduces to || x¢—r(t1) Il 2. More generally, if M= diag [m,,..., my]
then the m; are chosen so as to weight the relative importance of the
deviations [x; (t; )—7; (t1)] *. If some of the 7; (t, ) are not specified then
the corresponding elements of M will be zero and M will be only positive
semidefinite (r.s.p.s.d.).
Minimum effort. The desired final state is now to be attained with
minimum total expenditure of control effort. Suitable performance indices
to be minimized are
f‘ ulRudt (6.6
Optimal control 211
where R is r.s.p.d. and the 8; and 7; are weighting factors. We have already
encountered expressions of the type (6.6) in Chapter 4 (Theorem 4.4 and
Exercise 4.13).
Tracking problems. The aim here is to follow or ‘track’ as closely as
possible some desired state r(t) throughout the interval tg <t<tf,.
Following (6.4) and (6.6) a suitable index is
iPe'Qdet (6.7)
where Q is r.s.p.s.d.
ly
We introduced the term servomechanism for such
systems in Section 1.2, a regulator being the special case when r(f) is
constant or zero. If the u;(t) are unbounded then minimization of (6.7)
can lead to a control vector having infinite components. This is unaccept-
able for real life problems, so to restrict the total control effort a
combination of (6.6) and (6.7) can be used, giving
fe(iul>iar, (6.12)
212 Optimal control
<(x™Px)
LSEthe (6.1
whereP andQ satisfy the Liapunovmatrix equation (5.72), namely
A'P+PA=-Q. (6.17)
Optimal control 213
andintegrating
wehave
J, ={, "Oxdt
_
0
= { x™Pxdt—[tx7Px] : 0
=xeP, XotaOs
where
MP) +Pp AEP) 20) 1928.) (6.20)
Z + 2wtt + w?z=0
6? =(1+qw?)/4.
can be derived using the associated matrix equation (5.74). The perfor-
mance index corresponding to (6.15) is
xp QixK=-x} (APPA,—Pixg .
= x}Px,—xp41Px, poe (6.23)
Optimal control 215
Ko ye
= X09
Pxo,
provided A, is a convergent matrix so that x, > 0 as k > (Theorem
5.8). Similarly, multiplying both sides of (6.23) by k and summing leads
to
K,=> kf Ox,
k=0
=> xfPxy xdPro
0
a xo (P; —P)xo
with Py =P, and general expressions for the constants b,; can be derived.
Exercise 6.1.
In Example 6.1 show that if b > + and the missile is initially a distance
greater than c?/2(2b —1) from the aircraft then it cannot be caught.
Exercise 6.2.
If x(t) is the solution of (6.14), show by considering d[x(t) @x(t)] /dt
that if A is a stability matrix then
(es(x®x)dt=—B™'x9
@Xo,
where B=A @/J+/]®A.
216 Optimal control
Exercise 6.3.
Determine K, in the form (6.24).
Exercise 6.4.
Show that if xx is the solution of (6.21) and A, is a convergent matrix
then
co
This subject dates back to Newton, and we have room for only a brief
treatment. In particular we shall not mention the well known Euler equa-
tion approach which can be found in standard texts (e.g. Kirk 1970).
We consider the problem of minimizing J(u) in (6.13) subject to the
differential equations (6.1) and initial conditions (6.2). We assume that
there are no constraints on the control functions u;(t), and that J(u) is
differentiable, i.e., if uand u + 6u are two controls for which J is defined
then
AJ = J(u + §u)—J(u)
Integrating the last term on the right hand side of (6.26) by parts gives
Ja=Oberst] + J +p T¢+
Tx] dr px f
= 0g a)
[CPDagtS
6J,=|{——p'}6 -E
Lae
OFoH
tr et AT
|
MTXae, 6.29)
+ —6x+—but5x|dt,(6.29
where 5x is the variation in x in the differential equations (6.1) due to
du, using the notation
a =|me 0H
ax ae|
and similarly for 0¢/0x and 0H/du. Notice that since x(to ) is specified,
(5x)r=1, = 0. It is convenient to remove the term in (6.29) involving
5x by suitably choosing p, i.e. by taking
i adiees
mealtptle Rese, (6.30)
and
Pi(t1)=
od vt | (6.31)
218 Optimal control
f »Tip
Gtayaanen
tut) de (6.34)
subject to
X, =—ax,; tu, x,(0)=Xo (6.35)
where a and 7 are positive constants.
From (6.28)
H=xi +u? +p,(-ax, tu)
and (6.30) and (6.33) give respectively
pt =—2xt + apt (6.36)
and
2u* +/p;* =0, (6.37)
where xj and pj denote the state and adjoint variables for an optimal
solution. Substituting (6.37) into (6.35) produces
Xf =—axt—-tp}, (6.38)
and since ¢ = 0 the boundary condition (6.31) is just p; (7) = 0. The
equations (6.36) and (6.38) are linear and so can be solved using the
Optimal control 219
Pa ae aad ae ed ol
for tos et) s0r
i 1s r+(p)TS
ots FO")s
using (6.1). Since on an optimal trajectory (6.30) and (6.33) hold it
follows that dH/dt = 0 when u = u*, so that
(A)y=y* = constant, fo <t<4y. (6.39)
We have so far assumed that f, is fixed and x(t, ) is free. If this is
not necessarily the case, then by considering (6.27) the terms in 6J, in
(6.29) outside the integral are obtained to be
x*(t1) = xf (6.41)
(| og
=Fat eet (6.4
it,
In particular if ¢, F’, and f do not explicitly depend upon ¢ then (6.39)
and (6.42) imply
Resultant
velocity V
Current
FIG. 6.1
We see in Fig. 6.1 that the control variable is the angle u. The equations
of motion are
X, =Vcosut+c
= V(cos u—x2/a) (6.44)
x, =Vsinu, (6.45)
where (x, (f), x2 (t)) denotes the position of the ship at time t. The
performance index is (6.3) with tf) = 0, so from (6.28)
H=1+p,V(cosu-—x,/a)+p2V sinu. (6.46)
The condition (6.33) gives
—p*V sinu* + p*V cosu* =0
so that
tan u* =pF/pt . (6.47)
The adjoint equations (6.3u) are
pi =0,BF=prVia, (6.48)
222 Optimal control
Hence
(V/a)(dt/du*) = sec? u*
m[x*(t,)] =0 (6.53)
there are a further n conditions which can be written as
A
Pay oN ea ( ——
a deh?\ ——
ox heels Hedge—
\ ox 6.54
eae
x2
EG
aa OUTKeNt
as
FIG. 6.2
TABLE 6.1
t, fixed
Xp =Xs, 67 ae (6.57)
at time T so as to minimize
\e u? dt,
where c, and c, are constants. Substituting (6.58) and (6.59) into (6.57)
leads to
t t
xf=es6'—de,e'—deittcq, xf =—cse'— de,6° He)}
and the conditions
xT(0) = 0, xF(0) = 0, ax#(7) + bxF(T) =, (6.60)
must hold.
It is easy to verify that (6.54) produces
and (6.60) and (6.61) give four equations for the four unknown constants
cj. The optimal control u*(¢) is then obtained from (6.58) and (6.59).
In some problems the restriction on the total amount of control effort
which can be expended to carry out a required task may be expressed in
the form
so that
Xn +1 (t)=2(%, u, f). (6.63)
The differential equation (6.63) is simply added to the original set (6.1)
together with the conditions
Exercise 6.5.
A system is described by
x4 =—2x, aed
| 1 u? dt.
0
Show that the optimal control which transfers the system from x, (0) = 1
to x,(1) =O is u* =—4e?"/(e* —1).
Exercise 6.6.
The equations describing a production scheduling problem are
226 Optimal control
where /(t) is the level of inventory (or stock), S(t) is the rate of sales and
a is a positive constant. The production rate P(t) can be controlled and
is assumed unbounded. It is also assumed that the rate of production
costs is proportional to P*. It is required to choose the production rate
which will change /(0) = Jo, S(O) = So into (7) =J,, S(T) = S, ina fixed
time T whilst minimizing the total production cost. Show that the optimal
production rate has the form P* =k, + k2¢tand indicate how the constants -
k, and k, can be determined.
Exercise 6.7.
A particle of mass m moves on a smooth horizontal plane with rectangular
cartesian coordinates x and y. Initially the particle is at rest at the origin,
and a force of constant magnitude ma is applied to it so as to ensure that
after a fixed time 7 the particle is moving along a given fixed line parallel
to the x-axis with maximum speed. The angle u(t) made by the force with
the positive x direction is the control variable, and is unbounded. Show
that the optimal control is given by
tan u* = tan Uo + ct
where c is a constant and ug = u*(0). Hence deduce that
y*(t) = (a/c)(sec u* —sec ug)
and obtain a similar expression for X*(¢) (hint: change the independent
variable from f¢to wu).
Exercise 6.8.
For the system in Example 6.7 described by eqn (6.57), determine the
control which transfers it from x(0) = 0 to the line x, + 5x2 = 15 and
minimizes
In real life problems the control variables are usually subject to constraints
on their magnitudes, typically of the form | u;(t) | <k;. This implies that
the set of final states which can be achieved is restricted (see Section
5.7.1). Our aim here is to derive the necessary conditions for optimality
Optimal control 997
Example 6.8. (Takahashi et al. 1970, p. 637). Consider again the ‘soft
landing’ problem described in Example 6.2, where (6.12) is to be minimi-
zed subject to the system equations (6.9). The Hamiltonian (6.28) is
Since the admissible range of control is— 1 <u(t) < 1, it follows that H
will be minimized by the following:
u*(t) =—lifpx(t)>1
=(0if1>pk(t)>-1 (6.68)
=+1lifps(t)<-1.
Such a control is referred to in the literature by the graphic term bang—
zero—bang, since only maximum thrust is applied in a forward or reverse
direction; no intermediate nonzero values are used. If there is no period
in which u* is zero the control is called bang—bang. For example, a
racing-car driver approximates to bang—bang operation, since he tends
to use either full throttle or maximum braking when attempting to circuit
a track as quickly as possible.
In (6.68) u* switches in value according to the magnitude ofp¥(¢),
which is therefore termed the switching function, We must however use
physical considerations to determine an actual optimal control. Since
the landing vehicle begins with a downwards velocity at an altitude h,
logical sequences of control would seem to either
u* = 0, followed by u* = + 1
Consider the first possibility and suppose that u* switches from zero to
one at time f,. By virtue of (6.68) this sequence of control is possible if
pz decreases with time. It is easy to verify that the solution of (6.9)
subject to the initial conditions (6.10) is
xf=h-vt,xf=-v,0<t<t,
(6.70)
xt =h=vt+d(t-1,)?, xf =-v +(tt), t, <tST,
Substitutingthe softlandingrequirements(6.11)into (6.70)gives
T=hiv + dv, ty=h/v—4y. (6.71)
Optimal control 229
Because the final time is not specified and because of the form of A in
(6.67), eqn (6.43) holds, so in particular (H),,=,,* = 0 at t = 0, i.e. with
t = 0 in (6.67)
k + p¥(0)x7(0) = 0
or
pi(0) =k].
The adjoint equations (6.30) are
pt=0, Bf =—pt.
Hence
Di(t) =k/v,t>0
and
p3(t) =— kt/v-—1+kt,/v. (6.72)
using the assumption that p}(t, ) =— 1. Thus the assumed optimal control
will be valid if t; >0 and p3(0) < 1 (the latter conditions being necessary
since u*(0) = 0), and using (6.71) and (6.72) these conditions imply that
h>tv?,k<2v?/(h- $y?). (6.73)
If these inequalities do not hold then some different control strategy,
such as (6.69), becomes optimal. For example, if k is increased so that
the second inequality in (6.73) is violated then this means that more
emphasis is placed on the time to landing in the performance index
(6.12). It is therefore reasonable to expect this time would be reduced
by first accelerating downwards with u* = —1before coasting with
u* = 0, as in (6.69). It is interesting to note that provided (6.73) holds
then the total time 7 to landing in (6.71) is independent of k.
Example 6.9. Suppose that in the preceding example it is now required
to determine a control which achieves a soft landing in the least possible
time, starting with an arbitrary given initial state x(0) =x 9.The perfor-
mance index is just (6.3) with tp = 0, t; = 7. The Hamiltonian (6.28) is
now
H=\1 + P1X2 + p2u
MAS
7h1,p} <0;u*=-1,p, ao,
230 Optimal control
or more succintly,
u*(t) = sgn p2) (6.74)
Nie)
Hi OSs 7
a sc < tea
u*(t) = (6.75)
+105 fst trea
=1,0<1<t,;41,<7T<7
Integrating
thestateequations
(6.9)withu =+ 1gives
X, St Fl? +03f+04,x%2 =tt+c9. (6.76)
Eliminating
¢in(6.76)produces
x1(t)=3.x3(t)+s, whenu*=+ 1, (6.77)
x4(t) =—3.x3(t)+cg, whenu*=—
1. (6.78)
The trajectories (6.77) and (6.78) represent two families of parabolas,
shown in Fig. 6.3(a) and Fig. 6.3(b) respectively. The direction of the
arrows represents ¢increasing.
We can now investigate the various cases in (6.75):
X2 Xy
xX, xy
Cra OMCal)
P C—O
a0 ee (0
(b)
FIG. 6.3
(iii) | With the third case in (6.75), since u* =—1 fort; <t<T it
follows that x*(t, ) must lie on the curve QO. The point x*(f, )
is reached using u* = + 1 forO <t<f,, so the initial part of
the optimal trajectory must belong to the curves in Fig. 6.3(a).
The optimal trajectory will thus be as shown in Fig. 6.4. The
point R corresponds to ¢ =f, , and is where u* switches from
+ 1 to— 1; QO is therefore termed the switching curve. By
considering Fig. 6.3 it is clear that the situation just described
holds for any initial state lying to the left of both PO and QO.
FIG. 6.4
232 Optimal control
(iv) A similar argument shows that the last case in (6.75) applies
for any initial state lying to the right of PO and QO, a typical
optimal trajectory being shown in Fig. 6.5. The switching now
takes place on PO, so the complete switching curve is QOP,
shown in Fig. 6.6.
FIG. 6.5
FIG. 6.6
Optimal control 233
H=1+p'(Ax +Bu)
=1+p'Ax+(p'h,,... sp)DmJe
m
=1+pTAx
+s (p"bu, (6.80)
i=1
_ where the b; are the columns of B. Application of Theorem 6.3 to (6.80)
gives the necessary condition for optimality that
is the switching function for the ith variable. The adjoint equations
(6.30) are
pP=-
52(p*)"4x]
or
p* =—A'p*. (6.83)
Exercise 6.9.
A system is described by
d°z
dp ( )
me ee]AU
where z(t) denotes displacement. Starting from some given initial position
with given velocity and acceleration it is required to choose u(t), which
is constrained by | u(t) | <k, so as to make displacement, velocity, and
Optimal control 235
acceleration equal to zero in the least possible time. Show using Theorem
6.3 that the optimal control consists of u* = + k with zero, one, or two
switchings.
Exercise 6.10.
A linear system is described by
Z(t)+a2(t)+bz(t)=u(t)
where a > 0 and a* < 4b. The control variable is subject to | u(t) |<k
and is to be chosen so that the system reaches the state z(7) = 0, 2(7) =0
in minimum possible time. Show that the optimal control is u* = k sgnp(t),
where p(t) is a periodic function of f.
Exercise 6.11.
A rocket is ascending vertically above the earth, which is assumed flat. It
is also assumed that aerodynamic forces can be neglected, that the gravita-
tional attraction is constant, and that the thrust from the motor acts
_ vertically downwards. The equations of motion are
Ch
Eee af cu
oye dm_
gtin RU)
where A(f) is the vertical height, (¢) is the rocket mass, and c and g are
positive constants. The propellant mass flow can be controlled subject to
0 <u(t) <k. The mass, height, and velocity at time ¢ = 0 are all known
and it is required to maximize the height subsequently reached. Show
that the optimal control has the form
u*(t)=k,s>0; u*(t)=0, s<0
where the switching function s(t) satisfies the equation ds/dt =—c/m.
If switching occurs at time ¢, , show that
s(t)=+&nmea phat
Exercise 6.12.
A reservoir is assumed to have a constant cross-section, and the depth of
the water at time f is x, (t). The net inflow rate of water u(t) can be
236 Optimal control
xX =-0.1x, tu.
Find the control policy which maximizes the total amount of water in
the reservoir over 100 units of time, i.e.
{ 100 x, (t)dt.
0
iF 00u(t) dt = 60k
Exercise 6.13.
The following model can be thought of as representing a situation where
acup of hot coffee is to be cooled as quickly as possible using a finite
amount of cold milk, but could refer to a more general problem where
cold liquid is being added to a container of hot liquid. Suppose that the
coffee cup is originally full and at 100° C. The flow of milk, assumed to
be at a constant temperature, is represented by u(t) and is subject to
0 <u(t) <1. The total amount of milk is limited by
[pwde=1.
The thermodynamic equation describing the liquid in the cup is bilinear,
namely
e keNaas LOUa 1 UL
Nae
where the first term on the right hand side corresponds to heat loss to
the surroundings, the second term is due to the inflow of cold milk and
the third term is for the overflow. It is required to reduce the temperature
x, of the coffee to 0° C in minimum time 7. Show that the optimal
strategy isu* =0,0<t< 0.69; u* = 1, 0.69 <t < 1.69 (thus, somewhat
unexpectedly, there is a delay before the milk is added).
Optimal control 237
Exercise 6.14.
In Example 6.9 let x, (0) = a, x2 (0) = 6 be an arbitrary initial point to
the right of the switching curve in Fig. 6.5, with a > 0. Show that the
minimum time to the origin is 7* = B + (4a + 267)?.
Exercise 6.15.
For the problem in Example 6.8 suppose that the control is now to be
chosen so as to minimize the total fuel consumption, measured by
(e |u(t)|de
0
where 7; is fixed. Let the initial state be as in Exercise 6.14. If the
optimal control has the form
Th OCRALISfy
u*(t) = iis b<
eto: Sat Sel «
| showthat
t,t, = [Ti +BF KE SID —(40+267)}2];
and deduce that 7, = T* where 7* is the minimum time in Exercise 6.14.
Exercise 6.16.
Consider again the system described by the equations (6.9) subject to
| u(t) | <1.It is required to transfer the system to some point lying on
the perimeter of the square in the (x; , x2) plane having vertices (+ 1, + 1)
in minimum time, starting from an arbitrary point outside the square.
Determine the switching curves.
We have noted that a general closed form solution of the optimal control
problem is possible for a linear regulator with quadratic performance
index. Specifically, consider the time varying system (3.29), i.e.
with M and R(f) r.s.p.d. and Q(f) r.s.p.s.d. for t > 0 (the factors 4 enter
only for convenience). The quadratic term in u(t) in (6.86) ensures that
the total amount of control effort is restricted, so that the control variables
can be assumed unbounded.
The Hamiltonian (6.28) is
H=4x'Qx+4ulRutp! (Ax+Bu),
“)
a, [Eu* Rut + (p*)"Bu*] =Ru* +BTp*= 0,
so that
u* =—R™!BTp*, (6.87)
p* =—Qx*-A'p*. (6.88)
and combining this equation with (6.88) produces the system of 2n linear
equations
qf
dt | = A(t) — BR (QBY(t)][x*@
| . (6.89)
p*(t) SODcee Al) p*(t)
Since x(t, ) is not specified we have case (i) on p. 220, so the boundary
condition is (6.31) with ¢ = +x™Mx,ice.
aid @(t,ee))
t,) (01)
p*(t) 31 92]
| ie[x*(ty1 (6.91)
and substituting for X*, 5* from (6.89) and p* from (6.93) produces
(P+ PA —PBR7'B™P+Q+A"P)x*(t) =0.
Since this must hold throughout 0<t <¢;, it follows that P(t) satisfies
P=PBR~'B'’P—A™P-PA-Q (6.95)
where z(t) = J(t) w(2), v(£) is the input voltage, k, is a friction coefficient
and ky a torque coefficient. The speed of winding the wire on the reel
is s(t) = r(t) w(t), where 7(t) is the total radius at time ¢. It is required
to regulate the system so that s(t) is kept at a constant value so. Since
2(t)/s(t) = J(o)/r(t)
the corresponding value of z(t) is
Zo(t) = SoJ(t)/r(¢),
and from (6.97)
() 1 ( fe ty k1Zo
v0 cs aes
ie Be
0 J .
Clearly /(t) and 7(t) increase with time, and w(t) must be decreased in
order to keep s(t) constant.
Define as state variables
x1 (t) = 2(t)—Zo(t), x2(t)=s(t)—So
and as control variable
t +ku?)
Se(x3 dt,
Optimal control 241
u*(t) = —kyp(t)xi(O/k.
For a numerical case see the book by Kwakernaak and Sivan (1972,
p. 234).
It should be noted that even when the matrices A, B, Q, and R are
all time invariant the solution P(t) of (6.95), and hence the feedback
matrix in (6.94), will in general still be time varying. However of particu-
lar interest is the case when in addition the final time ft, in (6.86) tends
to infinity. There is then no need to include the terminal expression in
(6.86) since the aim is to make x(t, ) > 0 as t; >, so we set M= 0. Let
Q, be a matrix having the same rank as Q and such that Q = OF O07; plt
can then be shown (Kwakernaak and Sivan 1972, p. 237) that the solution
P(t) of (6.95) does become a constant matrix P, and we have:
THEOREM 6.5. If the constant linear system (6.79) is c.c. and [A, Q;]
is c.o. then the control which minimizes
is givenby
u* =—R™'B'Px, (6.99)
where P is the unique r.s.p.d. matrix which satisfies the so-called algebraic
Riccati equation
PBR7'B'P-A'P-PA-Q=0. (6.100)
Equation (6.100) represents tn(n + 1) quadratic equations for the
unknown elements of P, so the solution will not in general be unique.
However, it is easy to show that if a positive definite solution of (6.100)
exists (which is ensured by the conditions of Theorem 6.5) then there is
only one such solution (see Exercise 6.20).
The matrix Q,; can be interpreted by defining an output vector
y = Q;x and replacing the quadratic term involving the state in (6.98)
by y'y(= x7 QF Q;x).
242 Optimal control
X=Ax (6.101)
S™P+ Pg = ATP+PA-—2PBR=!B'P
(6.102)
=- PBR '*B'P—O,
using the fact that P is the solution of (6.100). Since R™is positive
definite and Q is positive semidefinite the matrix on the right in (6.102)
is negative definite, so by Theorem 5.17 & is a stability matrix. It can
also be shown (Kucera 1972) that if the triple [A, B, Q,] is not c.c. and
c.o., but is stabilizable and detectable (see Section 5.7.2) then eqn. (6.100)
has a unique positive semidefinite solution, and the closed loop system
(6.101) is asymptotically stable.
Thus solution of (6.100) leads to a stabilizing linear feedback control
(6.99) irrespective of whether or not the open loop system is stable. This
provides an alternative to the methods of Section 4.4. For example, in
the linearized model of the inverted pendulum problem presented in
Exercise 5.49, the system can be stabilized by linear feedback in all the
state variables, and this could be generated via (6.99) with a suitable
choice of elements of Q and R in (6.98).
If x*(£) is the solution of (6.101), then as in (6.16) eqn. (6.102)
implies
using(6.99).Since ®W
isa stabilitymatrix,wecanintegrateboth sides
of (6.103) with respect to t from zero to infinity to obtain the minimum
value of (6.98):
using the same argument as for (6.18). Of course when B = 0, (6.100) and
(6.104) reduce simply to (6.17) and (6.18) respectively.
A number of methods for solving (6.100) have been developed. One
Optimal control 243
and k is the feedback matrix b’P obtained from (6.99). It is easy to see
that (6.106) implies that the Nyquist locus of ¢(s) must not enter the
disc of unit radius centred on —1 + i0. It can be shown that the
gain margin (see Section 5.3) of the closed loop system is infinite and
that the phase margin is at least 60°. These are larger than necessary, and
this constitutes a practical disadvantage of using Theorem 6.5 for genera-
ting linear feedback.
For the multivariable case m > 1, eqns. (6.106) and (6.107) have been
generalized by MacFarlane (1970). It turns out that
A +t, Gico)| > 1, 7 = 1s i,
wherethe functions 1 + ¢,(s)are the characteristicroots of the optimal
return differencematrix
F(s) =Im +K(sl-—A) 1B
with K = R~!B"P, so the characteristic loci (see p. 172) must not enter
the unit disc centred on —1 + i0.
244 Optimal control
Exercise 6.17.
Use Theorem 6.5 to find the feedback control which minimizes
be(x2+0.1u?)
dt
subject to
x1 SN ait: tb X2 Chis
Exercise 6.18.
Use the Riccati equation formulation to determine the feedback control
for the system
Ka SaX4 tees
whichminimizes
S {1(3x?+u?)dt.
0
(Hint: in the Riccati equation for the problem put P(t) =~ w(t)/w(2)).
If the system is to be transferred to the origin from an arbitrary
initial state with the same performance index, use the calculus of
variations to determine the optimal control.
Exercise 6.19.
If the matrices X(t), Y(t) satisfy
blll
where L is defined in (6.105), show that P(t) = Y(t)X~1() is a solution
of eqn (6.95) with A, B, Q, R time invariant.
Optimal control 245
Exercise 6.20.
If P, and P, are two positive definite solutions of (6.100) show that
Hence deduce, using the uniqueness condition derived for eqn. (2.37),
that P, = °
Exercise 6.21.
Let y = Cx be an output-vector for the system (6.79), where C is a
constant matrix. If u is chosen so as to minimize
£ Pe(vyQy+u™
Ru) de,
where Q and R are both r.s.p.d., write down the appropriate Riccati
equation corresponding to (6.100) and denote its positive definite
solution by P,; (the system is assumed c.c. and c.0.).
Consider the dual system (4.25) with A, B, C time invariant, and
suppose that for this system wuis to be chosen so as to minimize
Bes(vIR7!y
+u™Q>1u)
dt.
Show that the positive definite solution of the associated Riccati equation
nPy.
Exercise 6.22.
Consider the optimal linear tracking problem in which it is required to
minimize (one half) the sum of (6.4) and (6.8) subject to (6.85). It can
be shown (Kirk 1970, p. 220) that eqn. (6.93) is replaced by
Show that the optimal control has the form u*(t) + v(t), where u*(f) is
given by (6.94) with P(t) satisfying (6.95) and (6.96), and
v(t) =—R7! BY s(t) where
ES |
FIG. 6.7
using the expression for f, in (6.109). Again using calculus of one variable
it is easy to show that
fa(c) = 3/27 (6.111)
with x, =c/3, and since c —x, = 2c/3 it follows by our result for n = 2
that x. =x3 =c/3. In view of (6.109) and (6.111) it is reasonable to
conjecture that
fn(c) = (e/ny" (6.112)
with
xX,=X_ =° °° =X, =Ce/n, (6.113)
LOTees tseo 1
248 Optimal control
sete Ut(pated)aoc
establishing (6.112). Also c -x, =cm/(m + 1) so by the induction
assumption x, =x3 =* * * =Xm =c/(m + 1), and this verifies (6.113)
In general it is not possible to obtain analytical results, as in the
preceding example. However, dynamic programming is essentially an
ingenious way of reducing a multistage decision process to a sequence of
single stage processes, and as such is a useful tool for constructing
algorithms for solving optimization problems. We now sketch an appli-
cation to a simple optimal control problem.
J [x(1)]=minJ12.
The penultimate stage k = 0 is then considered. By the P.O. the minimum
cost over the last two stages is
"
ic Far+ ae Far+o[x(¢1),t1]}
= =)
(6.116)
ein eng Fdr+J*[x(t
+Ad),
t+Ar]}
the last step following by the P.O. Assuming that J* can be expanded in
a Taylor series we have
A ttat oJ*
J* [x(t),
(x(t), t]t] = mi
— {{, F ar + 42J*[x(7),
tf] + pains
yi At
rN
ha
fia. [x(t+At)—x(t)]
+higher
order
terms
i (6.117)
Since J* is independent of u, the terms J*[x(f), t] cancel out in (6.117),
and if Af is small the resulting expression can be written as
oJ*
O=min(Farpo npr *earike terms
)'
u ot Ox
using (6.1). Dividing by Ar and letting At > 0 produces
*
ie = chant
GN: (6.118)
ice
250) Optimalcontrol
a* \ at
a(x,u,Zr) =r+ fi+ au*
aefie wG:1
Eqn (6.118) is a partial differential equation to be solved for J*, subject
to the boundary condition obtained by putting ¢ = ¢, in (6.116), namely
J*[x(t1), 01] = Oi), 41]. (6.120)
xi(T)+ |"‘e u? at
0
subject to the scalar equation
xy ae urile
The expression (6.119) gives
OH ee
y 0) ee GE
Substituting u* =— +dJ*/dx, into (6.121) gives
minH=4Ux,)?+Jz,(1—
2, )
u
=—403,)? txiJz,,
andeqn.(6.118)thenbecomes
(LM a 4a(2) .
ot Ox,
subjectto (6.120)whichisjust
J*[x(1), T] =xi(7).
In general such a partial differential equation must be solved numerically,
but in fact in this example an analytical solution is possible (see Exercise
6.24).
Optimal control 201
Exercise 6.23.
Use dynamic programming to find the numbers x;, x2, x3 which mini-
mize the sum x? + 2x3 + 3x3 subject tox, +x, +x3 =c, wherecisa
constant.
Exercise 6.24.
In Example 6.14 assume that J* takes the form k(t)xj, and hence obtain
the optimal control. Deduce that as T > © this approaches constant
linear feedback.
References
254 References
Chapter 1
12s 0 0) j 0) 0)
0 0 0 1 0
A= ; = ’
—k,/m, k,/m, 0 0 I/m,
Ky/m, —(ki +kz)/m, 0 0 0
123. Xy =X2,X2 = (3g/2) sin x, —3x3 sinx, cos x, —3X, sin 7x, +
(3M/m2)X4 cos x1,X3 =X4,
X4 =—(m/M) (Kq + WM.cos x, —Lx}?sin x,).
1.6. g182/(1 +182h —8183/).
1.7. x4(t) = [a2x2(0)—agx1(0) + €7(aix1 (0) —42x20) 4, d= ay ~ a4;
x,(t) =x, (0), x2(¢) = x2(0), all t > 0, irrespective of d.
1.9. gi(1 —g183h)/(1+g182h —g183h)’.
1.10. (@)Tz/(z-1)? (b) z/(z-e”*).
1.13. Fr =(0-096F>+ 0-85F,\(1-:063)*+ (0:906F'—0-85F;)(—0-1faye
My,=a, (1:063)*+a, (—0-113)*+45 (0-95),
a; constants.
Chapter2
sin 5+sin 1 sin 5 —sin 1
2.26. Allfinite A;sin A= + :
Sinvoasitin Sinvor-tasin:
Chapter3
Sealx,(t)
(@)|=(2x 1|+(x3(0)
,(0)-x2(0))e** 1
-s,0)e|
l 2
256 Answers to exercises
x(t) oak b
(b)|x2(t)|=@20)-Fx3(0)e?
| 1|+ 10) +x2(0))e*"]
-1
x3(t) 0 Ww
=a
+(x1(0)+x2(0)+ $x3(0)e**| 1
2
sinwt coswt
wes (1+de%! tert!
x(t) = x(0).
—te3! (1-fe734
=e ee ef —e2t e2t
3.16)Lpapiereat’ anaes
3 } Ty+ (2/9\(-1 +3t + &9*).
Laer MEPore?
3.20. ie = 510 (—210043100)
(218= 9.31%) 71 4.9.31) I
wot] =P 5= 23).
2) —67) +112) - 62 =u.
=3) = 6
3.31, eee
ae k Kia
Bie: be
ki
3:32 ' 1 1 a
ee C213 Oe Se
0 0 6
Chapter 4
4.2. c.c. if and only ifk,; #k.
4.4. 0 0 1 0
0 0 0 1
A=
~K/m, k,/m, —d,/m, d,/m,
ky /m, —(ky +k2)/my d,/m, —(d, +d)/m,
0
%* 0
| Amy
ky #4.
258 Answers to exercises
K= pees 371:
3 -4 1
4.38. —[33/2,14,6]!.
4.43, A=diag[s,,52,...,5p],B=[1,1,...,1]1,C=[ki,.-.,
kp).
4.46, Tegan!
[=r=| , 4.|
A=diag[1,—-1,-2],
B= ,C= 10/36 2 112
Chapter 5
51. (2,4).91=—
2y21+2),32=Vi1
V2+2).
53; kx? +x2 =c. 5.6. (a) Yes (b) No (c) No (d) Yes.
527. [exp(At)](x9-A™'b)+A‘ Dd.
535 One inside, three outside. RAY, KA 2.
52198 two.
FIG. Al
5.20 U unstable.
Sree — 2.
5.30. 1 13 ma!
P= 30 ; asymptotically stable.
=1 4
5.33. Asymptotically stable; +<<k<4;k>0.
a, 0 0 0
5.44. $x? +x3;x? + 2x3 <2. 5.46. stable. 5.47. asymptotically
stable.
5.48. No. 5.49. [0,—3/2,0, 4]! u/(4M +m); stabilizable by u = Kx.
5.54. uz = qjwj, qj (w) <0.
Chapter 6
6.3. xd(P—3P,+2P)xXo.
6.7. 4c (oasec u*+tan
+tan u*
ug)
+
?
| Sk Fhe
pee tem nobus
ey
a+—' ©,
a ial tard he
< rr
me
This is a concise, readable account of some basic
mathematical aspects of control. The approach
concentrates on so-called state-space methods, and
emphasizes points of mathematical interest. However,
undue abstraction has been avoided and the contents
should be accessible to those with a basic background in
calculus and matrix theory. The book will be useful to final-
year honours mathematics students, postgraduate control
engineers, and qualified engineers seeking an introduction
to contemporary control theory literature. Problems are
provided at the end of each section.