State Variables For Engineers - DeRusso, Paul M - (Paul Madden) Roy, Rob J - , Author Close, - 1965 - New York, Wiley - 9780471203803 - Anna's Archive
State Variables For Engineers - DeRusso, Paul M - (Paul Madden) Roy, Rob J - , Author Close, - 1965 - New York, Wiley - 9780471203803 - Anna's Archive
https://archive.org/details/statevariablesfoOOOOderu
-
State Variables for Engineers
State Variables for Engineers
PAUL M. DERUSSO ROB J. ROY CHARLES M. CLOSE
20 19 18 17 16 15
The feedback control field has been given a strong impetus in the
theoretical direction. Stability theory, as formulated by Lyapunov, and
modern optimization theory, as developed by Bellman and Pontryagin,
and the works of Kalman, LaSalle, Merriam, and others are responsible for
the changes. These works rely heavily upon “state variable” formulations,
and it is apparent that advanced presentations of control theory must be
given from this viewpoint. Furthermore, practicing control engineers
must learn this viewpoint, because most of the current technical papers
in this field utilize a state variable formulation. Otherwise, the current
divergence of theory and practice will increase.
Acquiring this body of knowledge is difficult, since most of the topics
that are published in book form are contained in advanced books on
mathematics. Many engineering students and practicing engineers do not
have the level of mathematical sophistication necessary to comprehend these
advanced treatments. The purpose of this book is to provide these people
with the necessary self-contained transitional presentation. It is intended
to follow a conventional first course in feedback control systems, and to
provide the necessary background for advanced presentations of nonlinear
theory, adaptive system theory, sampled-data theory, and optimization
theory. Furthermore, it attempts to unite the state variable approach and
the more usual transfer function concepts, so that the reader can relate
the new material to what he already knows. For these reasons, extensive,
complicated proofs are omitted in favor of numerous examples. We have
not tried to include all possible topics, but rather have attempted to cover
basic principles so that the reader can subsequently investigate the lit¬
erature for himself. Thus some treatments are not as advanced as can be
found in the literature. After completing this book, the reader requiring
more depth in state variables, Lyapunov theory, or optimization theory
is encouraged to read Reference 1 of Chapter 5, References 26 and 28
of Chapter 7, and Reference 25 of Chapter 8, respectively, and the recent
technical papers on these topics.
vi Preface
1 Time-Domain Techniques
1.1 Introduction 1
1.2 Classification of Systems 3
1.3 Resolution of Signals into Sets of Elementary Functions 8
1.4 The Singularity Functions 11
1.5 Resolution of a Continuous Signal into Singularity Functions 20
1.6 The Convolution Integral for Time-Invariant Systems 25
1.7 Superposition Integrals for Time-Varying Systems 32
1.8 Resolution of Discrete Signals into Sets of Elementary
Functions 35
1.9 Superposition Summations for Discrete Systems 37
2 Classical Techniques
2.1 Introduction 47
2.2 Representing Continuous Systems by Differential Equations 47
2.3 Reduction of Simultaneous Differential Equations 49
2.4 General Properties of Linear Differential Equations 52
2.5 Solution of First Order Differential Equations 53
2.6 Solution of Differential Equations with Constant Coefficients 54
2.7 Solution of Differential Equations with Time-Varying
Coefficients 68
2.8 Obtaining the Impulse Response from the Differential
Equation 71
2.9 Difference and Antidifference Operators 75
2.10 Representing Discrete Systems by Difference Equations 79
2.11 General Properties of Linear Difference Equations 82
2.12 Solution of Difference Equations with Constant Coefficients 83
2.13 Solution of Difference Equations with Time-Varying
Coefficients 94
vii
Contents
Transform Techniques
Index 603
State Variables for Engineers
1
Time-Domain Techniques
1.1 INTRODUCTION
r*~yi(t)
*~y2(t)
^y(t)
Vm(t) yn(t)
(a)
Fig. 1.1-1
2/(0 = ~t
at
v(0
The system is linear since
d d d
- [CjUiCO + c2v2(t )] = cx - vx{t) + c2 — v2{t)
at at at
2/(0 = v\t)
The system is nonlinear because
2/(0 = t - v{t)
cit
d d d
t — [cxVi(t) + C2V2(t) 3= Cit — Vi(t) 1
- - c2t — v2{t)
at at at
d „ , dv(t — X)
Example 1.2-5. Prove or disprove the following statement. In a linear system, if the
response to v(t) is y(t), then the response to Re [y(03 is Re [y(t)\. The symbol Re is read
“the real part of.”
Vi + Vx*
Re[y] =
2
where the asterisk superscript denotes the complex conjugate.
If 2/(0 = L[v(t)], the response to Re [y(/)j is
vx + vx*
L
2 — \L[V i] +
Example 1.2-6. Prove or disprove the following statement. In a linear system, if the
response to v(t) is y(t), the response to (d/dt)v(t) is (d/dt)y(t).
If 2/(0 = L[v(t)\, the response to dvjdt is
dv
L — L[v\, L not a function of t
dt dt
Example 1.2-7. Prove or disprove the following statement. The cascade connection of
two linear systems, as shown in Fig. 1.2-1, is linear.
<7(0
v(t) ■> >- y(0
Fig. 1.2-1
q(t) = Lx[v(t)\
Since the two component systems are linear, the responses to v(t) = c1y1(/) + c2vft)
are given by
q(t) = Lx[cxvx(t) + c2v2(t)] = c1L1[u1(t)] + c2Lx[v2(t)]
y(t) = L2{cxLx[vx{t)\ + c2Lx[v2(t)]}
= cxL2{Lx[ux(t)]} + c2L2{Lx[v2{t)}}
6 Time-Domain Techniques
By the definition of linearity, the last equation proves that the cascaded combination
of N1 and N2 forms a linear system. It is true, in fact, that any system composed only
of linear components is itself linear.
fi(t) h(t)
h(t)
Fig. 1.2-2
Example 1.2-9. Continuous signals that change only at discrete instants may be
completely described by equivalent discrete signals. Figure 1.2-3a shows a staircase
fi(t) hit)
->- t 1
o T 2T
(a) (b)
h(t) MO
1
0 T 2T
(c) (d)
Fig. 1.2-3
function whose value changes only at / = kT(k = 0, 1, 2, . . .), and Fig. 1.2-3& shows
a series of pulses of known width. These continuous signals are equivalent to the
discrete signals in parts (c) and (d) of the figure, respectively, in the sense that the original
signals can ideally be regenerated.
Example 1.2-11. In pulse code modulation, each time increment of signal is quantized,
often to the nearest of eight levels, including zero. Each quantized increment is then
represented by three two-level pulses using the binary number system. Thus the signal
of Fig. 1.2-4a is quantized and coded as shown in parts (b) and (c), respectively.
The next chapter discusses the classical solution for the response y(t)
to a given input v(t). Depending upon the nature of both the system and
Sec. 1.3 Resolution of Signals into Sets of Elementary Functions 9
the input, classical techniques may involve considerable effort. One would
certainly expect, and it is true, that the responses to certain classes of
input functions could be determined more easily than the response to an
arbitrary input. Accordingly, a sensible procedure for linear systems is
to try to express a given arbitrary input as the sum of elementary functions.
If the response to each of the elementary functions is known, or if it can
be easily found, then the response to the arbitrary function can be found
by the superposition principle of Eq. 1.2-2.
Although this scheme is satisfactory for linear systems, whether they
are time-varying or not, it cannot be extended to nonlinear systenls. Its
validity is based upon Eq. 1.2-2, which does not apply to nonlinear
systems. This section describes the scheme in general terms, and the next
few sections and parts of Chapters 2 and 3 deal with the important special
cases. The reader may wish to return to this section later.
Computational ease is not the only reason behind the proposed scheme.
The characteristics of a system may be expressed in several different ways,
e.g., by differential equations, or by the response to certain elementary
functions. Although it is usually possible to convert from one method of
characterization to another, it is not always easy to do so. Furthermore,
different methods of characterization give different insights into system
analysis, and certain problems are much more conveniently treated by a
particular approach, such as the one discussed here.
Suppose that the input functions under consideration can be decom¬
posed into a denumerable number of elementary functions denoted as
k/f = 0, 1,2,...). Then
00
v(t) = 2 ajilt)
i=1
y(t) = ^ a(X)K(t, X)
A
(1.3-2)
As discussed in Chapter 3, the Fourier series for a periodic function has
the form of Eq. 1.3-1. For nonperiodic functions, however, the resolution
of v(t) into a denumerable set of elementary functions is not possible. A
continuous set of elementary functions is then required. k(t, X) and a(X)
become continuous functions of the parameter X, and Eq. 1.3-1 is replaced
by
a(X)k(t, X) dX (1.3-3)
J
k(t, X) and a(X) are still called the elementary and spectral functions,
respectively. If the response of a linear system to k(t, 2) is again denoted
by K(t, X), the response to v(t) is
(i) There must be an easy way of finding the coefficients a(X) for any
arbitrary function of time.
(ii) There must be an easy way of finding K(t, X), the response to the
elementary function, for an arbitrary linear system.
The singularity functions are a class of functions that forms the basis
for the time-domain analysis of systems. Every singularity function and
all its derivatives are continuous functions of time for all real values of t
except one. Furthermore, the singularity functions can be obtained from
one another by successive differentiation or integration. The singularity
function that is perhaps the most familiar is the unit step function, which
is denoted by U_L and is shown in Fig. 1.4-1. As can be seen from the
figure,
0 for t < 0
(1.4-1)
1 for t > 0
U_x(t) U_1(t - a)
k /
- 1
r-H
H
1
1
1 >
0 0 a
(a) (b)
Fig. 1.4-1
(0 for t < 0
U_2(t) = U_.it) dt = (1.4-3)
I —CO it for t > 0
and is shown in Fig. 1.4-2a. Another singularity function is
rt fo for t < 0
U_3(t) = U_2(t) dt = (1.4-4)
-oo for t > 0
and is shown in Fig. 1.4-26. The subscript notation is suggested by the
following symbolism, sometimes used in mathematics.
f~\t) = j7(0 dt
r\t)=jr\t) dt
U_ 3 (t)
Fig. 1.4-2
Sec. 1.4 The Singularity Functions 13
fo(t)
(a) (b)
Fig. 1.4-3
m = y/_,(»)
dt
(1.4-5)
/-i(o = f m dt
J — 00
lim/_1(0 = CLjO)
A -* 0
rT / \ r. fO for t t6 0
U0(t)= hm/0(0 = _ _
a-*o unhmty lor t = 0
U0(t) is called the unit impulse function, and its properties are shown by
the representation of Fig. 1.4-4. The number 1 alongside the arrow is
14 Time-Domain Techniques
Uo(t - a)
Uo(t): oo
A
A A
l l
1 t
0 0 a
(a) (b)
Fig. 1.4-4
intended to infer that the total area underneath the “curve” is unity. In
the limit, Eq. 1.4-5 becomes
t7_x(0 = lim
A -> 0
U0(t) = y [/_!»
dt
(1.4-6)
u_i(») (70(t) dt
This last step does not, however, necessarily follow by rigorous mathe¬
matics. Differentiation and integration are both limiting processes
themselves, and Eq. 1.4-6 is valid only if the differentiation and integration
are commutative with taking the limit as A approaches zero. That
interchanging the order of two limiting processes is not always valid can
be seen by noting that
_ y^
lim lim —-— ^ lim lim—-—
x~*0 y~*0 y y-*0 x~*0 X“ -f- y~
always, the set of all real numbers. The definition above makes the terms
function and single-valued function identical. A “multivalued function”
may be represented by two or more single-valued functions.
The unit step function is a perfectly good mathematical function. If one
starts to define the unit impulse function by the relationship
this is satisfactory also. But the definition of the unit impulse by taking
the limit off0(t), shown in Fig. 1.4-36, must include the fact that the area
under the curve is unity. This can be indicated by
f U0(t) dt = U_.it)
J — 00
or, equivalently,
U0(t — r) = 0 for t 5^ t
%t
U0(t — r) dt = U_.it - r) (1.4-7)
J — CO
00
The first two properties have already been discussed. The third can be
proved by noting that, since the integrand is zero except at t = r,
00 'r+e
f(t)U0(t - t) dt = f(t)U0(t - t) dt
-00 ’T—e
f The standard work on this subject is Reference 2. A simpler and more readable
treatment is given in Reference 3.
16 Time-Domain Techniques
When a unit impulse is the input to a system with energy storing ele¬
ments, it serves to change the system’s stored energy instantaneously.
If a unit impulse of force is applied to a fixed mass M that has previously
been at rest, the velocity of the mass is
i
e(t) = —
r 1
Uo(t)dt — — U_i(t) volts
C J—oo C
o-+
A
CeoUo(t - to)
amperes eoU-, (t - to)
volts
io^-1 (t - to)
amperes
fo(0 fo(t)
fi(t)
A
Ui(t): oo
/
A2
1
A 2A ->■1
o
0
1
A2 (
— oo
(c)
(d)
Fig. 1.4-7
The limit of the rectangular pulse of Fig. 1.4-36 is not the only way of
arriving at the unit impulse. Any function which possesses the desired
properties in the limit may be used in place of a rectangular pulse. Two
such functions are shown in Figs. \A-la and 1.4-76. In each case, as
A approaches zero, the unit impulse results. The derivative of the function
in part (b) of the figure is shown in part (c). In the limit as A approaches
zero, f0(t) becomes the unit impulse, and fx(t) becomes the unit doublet
of part (d). The unit doublet has the characteristics
(0 for t 0
l/i(0 =
— oo and + oo for t = 0
(1.4-8)
U0(t) = U^t) dt
' — 00
The singularity functions that are the most useful as elementary functions
are the unit step and the unit impulse.
Later sections of this chapter show how the response of a linear system
to any arbitrary input can be determined, once the response to a unit
impulse or unit step function is known. The terms impulse and step response
are always understood to mean the response to unit singularity functions
when the system contains no stored energy prior to the application of the
input.
Example 1.4-1. If the input to Fig. 1.4-8 is i\(t) = U0(t), it acts like an open circuit
except at t = 0, when it instantaneously inserts some energy into the circuit. Because
of the presence of the inductance, the impulse of current will all flow through the right-
hand resistance, causing an impulse in the response e0(t). Since the current through the
RL branch will remain finite, the impulse of voltage will appear directly across the
inductance, creating a current of \ ampere. For t > 0, the circuit may be redrawn as
in Fig. 1.4-9, where /(0+) = £. The circuit has a time constant of 1 second, so
Since the circuit is time-invariant, the step response is the integral of the impulse
response, namely,
eo (0
Fig. 1.4-8
20 Time-Domain Techniques
Fig. 1.4-9
For most systems, the impulse and step responses are determined
analytically from the differential equations describing the system. The
relationship of the impulse response to the system’s differential equation
is discussed in detail in Chapter 2. Its relationship to the system function
of Laplace transform theory is covered in Chapter 3. In some cases, the
impulse or step response may be given or may be approximately deter¬
mined experimentally.
Consider the function v(t) shown by the solid curve in Fig. 1.5-\a.
It can be crudely approximated by the staircase function, shown by the
broken line. The staircase function is the superposition of the five step
functions in part (b) of the figure.
where Av(k AA) denotes the jump at the point k AA in the staircase
approximation of Fig. 1.5-26. Since the factor U_x{t — k AA) = 0 for
t < k AA, it ensures that the jumps in the staircase function to the right
22 Time-Domain Techniques
v(\)
of point t have no bearing on the value of v(X) at the point t. Note also
that
dv(X)
Ar>(/c AA) === AA
L dX J k AA
Thus
00
dv(X)
v(t)= 2 C/_x(r - k AA) AA (1.5-1)
fc=—oo l^Tj fc AA
All three of the approximations above become exact as AA approaches
zero. In the limit, AA becomes dX, k AA becomes the continuous variable
A, and the summation is replaced by an integration.
oo
dv(X)
v(t) = U_t(t - X) dX
dX
where the quantity [v(fc AA)] AA is the area of the rectangular pulse
beginning at the point k AA in Fig. 1.5-26. The approximation becomes
exact as AA approaches zero, yielding
r(t, A) dX (1.5-5)
dX
Next let the input be approximated by Eq. 1.5-3. Denote the response
of the system to a unit impulse occurring at t = f by the symbol h(t, ft).
By superposition, the approximate response to an arbitrary input is
oo
In many problems, the input is zero for negative values of time, in which
case the lower limit for the integrals in Eqs. 1.5-5 through 1.5-8 becomes
zero.
24 Time-Domain Techniques
v(t)
(b)
Fig. 1.5-3
There are two situations which sometimes cause difficulty in the use of
these equations. If the input v(t) has discontinuities, its derivative, which
appears in the integrand of Eq. 1.5-6, will contain impulses. In the
evaluation of the integral, the sampling property of Eqs. 1.4-7 and 1.5-4
is useful. The derivative of the input shown in Fig. 1.5-3# has impulses
of value A and B, respectively, at t = 0 and /0. Thus Eq. 1.5-6 reduces to
where dvJdX stands for dvjdX with the impulses removed. Their effect
is accounted for by separate terms. The last equation follows directly
from the sampling property of impulses, or alternatively from breaking
the function of Fig. 1.5-3# up into the three components shown in part (b).
The impulse response of a system may occasionally contain an impulse
itself. Then the integrand of Eq. 1.5-8 contains an impulse. In this case,
Sec. 1.6 The Convolution Integral for Time-Invariant Systems 25
The preceding derivations are valid for all linear systems. For time-
invariant systems, a further simplification is possible. The symbols h(t)
and r(t) are customarily used to stand for the responses to a unit impulse
and unit step, respectively, applied at t = 0.
Equations 1.5-5 through 1.5-8 then lead to Eqs. 1.6-2 and 1.6-3 in Table
1.6-1. Equations 1.6-4 and 1.6-5 in the table are easily obtained from the
first two by the substitution of variable r = t — A. For nonanticipatory
Table 1.6-1
% OO
y(t) = v(X)hC - X) dX ' G-6-2)
J — OO
r
Upper limit is t Lower limit is 0
y(t) = (1.6-3)
l — oo
r oo
y(0 = v(t — r)h(r) dr G-6-4)
' — 00
Lower limit is 0 Upper limit is t
systems with v = 0 for t < 0, the limits on all the integrals become 0 and t.
Even in this case, however, it is not incorrect to use the wider limits, and
this is occasionally done to facilitate the proofs of some theorems.
Equations E6-4 and 1.6-5 suggest another interpretation for time-
domain analysis. Consider Eq. 1.6-4 for a nonanticipatory system with an
input that is zero for t < 0.
y(t) and v(t) represent the output and input, respectively, at the instant of
time t. Since v(t — r) represents the input r seconds before this instant,
r is sometimes called the age variable. As r increases, v(t — r) represents
the input further and further back into the past. For r = 0 and t, re¬
spectively, v(t — r) is the input at the instants t and 0, respectively. The
equation above indicates that the entire past history of the input con¬
tributes to the output at the instant t. The past history is weighted,
however, by the factor h(r). In fact, h{r) is sometimes called the system¬
weighting function instead of the impulse response. This interpretation
in terms of a weighting function will, perhaps, become clearer later when
the graphical interpretation of the equation is considered. For stable
systems, h(r) approaches zero as r approaches infinity. For such systems,
the more recent past is weighted more heavily than the far distant past.
Equations 1.6-2 and 1.6-4 may be rewritten in still another way if
desired. Simply use the fact that, for time-invariant linear systems,
=
Jo
fViWUt - X) dX (1.6-7)
Example 1.6-1. For the circuit and voltage source shown in Fig. 1.6-1, find the
current i(t). Assume that the circuit has no stored energy for t < 0.
In the preceding terminology, v(t) = e(r) and y(t) = /(/). The current response to
a step of voltage is
The equation
de{X)
i{t) = r(t - X) dX
dX
will be used. Since the character of the input is different for 0 < t < 1 and for 1 < t,
the solution must be carried out in two parts. For 0 < t < 1,
t--1(1 - e~{R'L)t)
R
For 1 < t,
KO (1) — [I €-UtlL)U-X)]dX +
R
1
\ --W - 1)«
R R
Note that the two solutions are identical for t = 1, as expected. Also note that the second
solution is not the first solution with t replaced by 1.
Example 1.6-2. Figure 1.6-2 shows a cascade connection of two identical circuits.
Assuming that the impulse response of each one individually is h{t) — te~t for / > 0,
and that the second circuit does not load down the first, find the impulse response of
the entire combination.
Fig. 1.6-2
28 Time-Domain Techniques
When
(0 = U0(t)
then
e*(0 = te~l for / > 0
and
t
e*(t) = e2(X)h(t - ;.) dX
o
rt 3, —<
The last expression represents the impulse response of the entire combination.
Example 1.6-3. Assume that the individual circuits of Fig. 1.6-2 are amplifiers, each
having the step response r(t) = — A€~t,RC for t > 0. Find the step response for the
two stages in cascade, assuming that the second does not load down the first. Extend
the result to n stages.
When
= tf-i (0
then
e2(t) = —Ae~t/Rc for t > 0
and
de2(X)
r{t — X) dX
dX
de2(X)
= e2(0 )r(t) + r{t — X) dX
'o+ dX
The last expression has a separate term as a result of the discontinuity in e2(t) at t — 0.
This is similar to Eq. 1.5-9.
P ’ A r-X!RC
«,(/) = +
io RC
_ ^2e~t!RC
1 - for t > 0
RC
The last expression is the step response for two stages in cascade. For three stages in
cascade, the step response is
—t!RC\
f4 A2 —d €~A/rO
/j A \
[-A€-{i-WRC]dX
L l RC) J
O
(n — 1)! (—tlRC)k
(~A)ne-tlRC 1 + > --—---- for t > 0
K Zv (n-k — 1)!(A!)2
A:=l
The last expression can be more easily obtained by means of the Laplace transform,
and it is derived in Chapter 3. The expression has been plotted for various values
of nd
Sec. 1.6 The Convolution Integral for Time-Invariant Systems 29
m *m = I7
Jo
i(A)/2« - ^ dx (1.6-8)
is based on the fact that the integral of a function between two limits
represents the area under a curve between those limits. For concreteness,
let
for 0 < t < 1
/i(0 =
elsewhere
m = 711 - e~(R,L)t]
XV
for t > 0
fiW
tV
1
Fig. 1.6-3
30 Time-Domain Techniques
v(t) h(t)
mirror image of the original function about the vertical axis. In plotting
this quantity versus X, it is shifted t units to the right. The multiplication
of/i(T) and f2(t — X) produces the last curve in Fig. 1.6-3. The shaded
area represents the integral in Eq. 1.6-8 for a typical value of t. In
summary, to find fx{t) * fold one of the functions about the vertical
axis, slide it a distance t into the other function, and take the area under¬
neath the product curve.
Graphically determining this area for different values of t will lead to a
plot of fY(t) * f2(t) versus t. In the example above, both the analytical
and the graphical approaches indicate that the response monotonically
increases from 0 at t = 0 to l/R as t —► oo, but that the rate of increase
suddenly decreases at t — 1. Note that the result would not be changed
if the limits of integration were changed to — oo and + oo. The integrand
is zero except for 0 < X < t, since fi(X) = 0 for X < 0, and f2(t — X) = 0
for X > t. For anticipatory systems with inputs for t < 0, the limits
would have to be — oo and + oo.
The graphical evaluation of the convolution integrals is particularly
useful when the analytic evaluation proves to be difficult, or when the
input or impulse response is given graphically instead of analytically. It
also reinforces the previous interpretation of the impulse response as a
weighting function. Consider the formula
r*t f*t
y{t) = v(X)h(t — X) dX = v(t — r)/l(r) dr
Jo Jo
Typical v and h functions are shown in Fig. 1.6-4. v(t — r) represents the
input t seconds before the “present” time t. Thus t = 0 corresponds to
the present; r > 0, to past time; and t < 0, to future time, as indicated
in the figure. Since v(t — r) is multiplied by h(r), the past values of the
input are weighted by the impulse response.
Sec. 1.6 The Convolution Integral for Time-Invariant Systems 31
The convolution of two integrable functions fx(i) and f2(t) was given by
oo
/i(0*/2(0= A(t - X)f2(X) dX
-oo
oo
= AWUt - A) dX (1.6-9)
' — oo
where the more general limits of — oo and -f oo have been used. Differ¬
entiating both formulas with respect to t, it is found that fff) * ff{t) =
fi(t) */2(0, where the prime denotes differentiation. In general,
m * f?\t) = f\kXt) * m
(1.6-10)
m ftkxo f[~kXt) * m
* =
where the superscripts (k) and (—k) denote the kth derivative and integral,
respectively.5 The first of these two expressions is written out explicitly
below.
00 oo
•(k)
/i(A)/?’(r - A) dX = I f[kXX)Ut - X) dX (1.6-11)
-oc -oo
where
/lW(A) = /l(A)
dXk
d>
AkX‘ - A) =
d(t - Xf
Mt - A)
The use of Eq. 1.6-11 for k = 1, together with the fact that h{t) = (dr/dt),
enables Eqs. 1.6-3 and 1.6-5 to be derived from Eqs. 1.6-2 and 1.6-4,
respectively. Equation 1.6-11 can also be helpful in the evaluation of the
convolution integrals, whether using analytical, graphical, or computer
techniques.5 Another application is the generalization of the impulse
sampling property of Eq. 1.5-4 to higher-order singularity functions.
Letting f2(t) = U0(t), Eqs. 1.6-11 and 1.5-4 yield
oo
dkA(t)
MX)uk(t - X) dX = for k = 0, 1, 2, . . (1.6-12)
-00 dtk
A modification of Eq. 1.6-11 is used in Section 2.8 to relate a system’s
impulse response to its differential equation. Note that
In general,
d%(t - X)
[/»(< - A)] = (-1)'
dXk d(t - Xf
Equation 1.6-11 can then be written
r.
J — oo
AW
dk
_dXk
Mt - X) dX = (-If
00
—oo _d A
.AW [A(t A)] dl
(1.6-13)
For the special case where f2(t) = U0(t), this becomes
oo
r d*
AW C/0(t - A) dA = (-1)*^© (1.6-14)
' — 00 dA/c dtk
The formulas of Section 1.5 are valid for all linear systems, whether
they be fixed or varying. The formulas of Section 1.6 are based upon
Eq. 1.6-1, which is valid only for fixed systems. The analogous results
are now established for varying systems.
Equation 1.6-1 indicated that for fixed systems the impulse and step
responses are a function only of r, the time that has elapsed since the
application of the impulse or step function. For varying systems, these
responses are functions of two variables. These may be taken as the present
time t and the time A at which the singularity function was applied or,
alternatively, the present time t and the elapsed time r. It must be clearly
understood which pair of variables is being used. Thus h(t, p) could be
interpreted as the response at time t resulting from an impulse applied
at either time p or time t — p. Since both interpretations are used in the
literature, this causes some confusion. In this book, the symbol /?(/, /3)
is always given the first of the two interpretations. When the second
interpretation is intended, the symbol h*(t, p) is used.f Specifically,
Equations 1.5-5 through 1.5-8 lead to the results in the first half of Table
1.7-1. The last half of the table is obtained by the substitution of variable
t = t — X. In Eqs. 1.7-2 and 1.7-3 the h(t, 2) and r(t, X) functions are more
commonly used, while in Eqs. 1.7-4 and 1.7-5 the h*(t, r) and r*(7, t) func¬
tions lead to more compact results. Since the term convolution is restricted
to formulas like those in Table 1.6-1, the equations in Table 1.7-1 are
Table 1.7-1
Simplification
When System Is Simplification when Equation
General Equation Nonanticipatory v = 0 for t < 0 Number
r oo
2/(0 = v(X)h{t,X)dX
J — 00
0-7-2)
f* 00
2/(0 = v(X)h*(t, t — 7) dX
J— oo
Upper limit is t Lower limit is 0
OO
dv(X)
2/(0 = — r(t, X) dX
— co
dX
(1.7-3)
OO
dv{X)
2/(0 = r*(t, t — X) dX
• 00 dX
00
y{t) = v(t — r)h(t, t —t) dr
' — 00
(1.7-4)
00
y(t) = v(t — r)h*(t,T) dr
I— oo
Lower limit is 0 Upper limit is t
00
dv(t — r)
y(t) = r(/, t — t) dr
d(t — r)
J — 00
(1.7-5)
00
du{t — r)
2/(0 r*(t, r)
1— 00 £/(t — t)
usually called superposition integrals. Table 1.6-1 for fixed systems can, of
course, be regarded as a special case of Table 1.7-1. For fixed systems,
h(t, X) reduces to h(t — X), and h*(t, r) to h{r).
h*(t, t) can be interpreted as a weighting function for varying systems,
exactly corresponding to the weighting function h(j) for fixed systems.
The graphical interpretation of the convolution integrals can also be
extended to the general superposition integrals. The discussion associated
with Fig. 1.6-4 applies to varying systems if hfr) is replaced by h*(t, r).
To find
2/(0 = I v(t — t)/i*0, t) dr
Jo
34 Time-Domain Techniques
fold v(t) about the vertical axis, and slide it forward a distance t to form
v(t — t) versus r. Multiply the resulting curve by h*(t, r), and take the
area underneath the product curve. The only difference from the fixed
system procedure is that, since h*(t, r) is a function of time, a different
h*(t, t) curve must be used for each value of t considered. The adjoint
technique of Chapter 5 proves helpful when using h*{t, r) for a fixed t
and varying r.
Equations 1.6-9 and 1.6-11 are mathematical identities that may be used
whenever the integrands have the proper form. One of the functions must
be an explicit function of 7, and the other an explicit function of t — 7,
where t is treated as a parameter. Starting with Eq. 1.7-5
dv(t — r)
r*(t, t) dr
d(t — r)
Eq. 1.6-11 gives
where the lower limit becomes zero for nonanticipatory systems. Also
By Eq. 1.7-2,
y(t) =
Sec. 1.8 Resolution of Discrete Signals into Sets of Elementary Functions 35
The same example is solved in Example 3.9-3 using the Laplace transform.
The only discrete signals considered in this book are those which are
defined at equally spaced instants, denoted by t = tQ + kT. In this
chapter, k is a time index which takes on only integral values, and T is
the time between instants. While t0 may have any value, the time origin
may be chosen to make t0 = 0, and this is consistently done. It is also
possible to choose the time scale so that T = 1, as in Sections 2.9 through
2.12, but this is not done in this chapter.
A discrete signal consisting of a series of finite numbers may arise
naturally, as in the case of a digital computer, or it may result from
sampling a continuous function, as in Figs. 1.2-2c and 1.8-lc. If the ideal
periodic switch in Fig. 1.8-16 is assumed to close every T seconds for an
infinitesimal time,/2(r) is zero except at the sampling instants. In terms
of the unit delta function
(1 for / = kT
S(t — kT) = (1.8-1)
|p for t t6 kT
the signal f2(t), shown in part (c) of the figure, can be expressed as
00 oo
Consider the more practical situation where the ideal switch in Fig.
1.8-16 remains closed for AT seconds. The signal f2(t), shown in part (d)
of the figure, is a continuous signal that may be approximated by a dis¬
crete signal. As discussed in Sections 1.4 and 1.5, a narrow pulse may be
approximated by an impulse of equal area, provided that the time con¬
stants of the system which follows are large compared to the pulse width.
The signal of Fig. 1.8-1 d can be approximated by
00
flit)
flit) h(t)
—-—>
(b)
h(t) h(t) \
>
—N
\
X
\ \
\
\ \
\
\ \
\ \
\
\
\
\
\
\ \
\
,<rrl
1
0 T 2T 3T AT 5 T 6T 0 T 2T c T AT bT 6T
(c) (d)
OO oo oo °o oo oo
v / v /k /^ I k A
1 1 1 1 1 1
t
0 T 2T 3T AT 5T 6T
(f)
>
_is
o
o
/
/
0 ^
CO
\
8^-
\
\
\
\
0 T 2T 3T AT bT bT
(h)
(g)
Fig. 1.8-1
where 2 U0(t — kT) represents the impulse train shown in part (/). The
k=— oo
idealized sampling switch of part (h), together with the symbol fx*(/),
is defined by Eq. 1.8-4. Such a switch modulates the input by a train
of unit impulses to produce a discrete output signal composed of impulses
of varying area. This is the sampling device assumed in books on sampled-
data control systems, and it is useful in later parts of this book. The
degree of approximation involved in going from part id) to part (g) of
Fig. 1.8-1 is best discussed in terms of transform techniques.6 It should
be noted that all of the ideal switches discussed in this section are linear,
time-varying components.
If the input to a linear system initially at rest is the discrete signal given
in either Eq. 1.8-2 or 1.8-4, the output can be found by superposition.
In this section, any sampling switch needed to produce the discrete input
is not considered to be a part of the system under consideration. The only
information needed about the system is its response to a unit delta or
impulse function, whichever is appropriate.
Time-Invariant Systems
The response of a fixed system depends only upon the time that has
elapsed since the application of the input. Let h{t) and d(t) denote the
response to a unit impulse and delta function, respectively, applied at
t = 0. Then the response to U0(t — kT) or to d(t — kT) is h(t — kT)
or d(t — kT), respectively.
Except in trivial cases, the delta response d(t) is identically zero for
oo
continuous systems. Any input for which |y(l)| dt = 0 does not insert
J — oo
any energy into the system and hence cannot cause an output to occur.
In discrete systems, such as a digital computer, a finite discrete input is
converted to a finite discrete output. In such a case, d(t) is defined only
at the instants / = t0 + nT, where n is an integer. The term t0 represents
a possible constant time delay between the input and output sequences.
Since the chief concern is with the form of the output, t0 will be ignored
38 Time-Domain Techniques
C
O Vh
5 s /.—N
G JO
<?i G <N m Tf VOI
t
G % On On o\ 6) Os
cr ^ i
Os
* h^H —i
w Z
G on o
O
•
~ V +-> -*—> -t->
3
oj
o Oh
c Vi 1 O E *|| .1 o £ O
-3 Uh •S —< l_ rn Ui
£ 1 a>
Uh «
Vi
a<i>
£X
G
C
o
J-h
<u
£
^
.23
£ ^
Dh 00
<[>
^
N
oo
®N
V•
—i
Vh
<D
£
N
C/5
a>
Dh .£
^
<l> o Dh o Oh
00 ^ 3“ o
N G P hJ
w H-J P
V5
3
c •- &■ <D
.2 g 2 ■*—»
"
™ -2 S « E +->
O oo a, o i II
3
00 E R
O
G >>. o Lh ||
33 00 '-2 «i -V
<o
N I-H
hi
u
i-H
<u
Dh - C <U <u a> £ ^ a> N
.§1 g Oh
Dh’~<
00
o
£ C/5 Dh
Dh
Dh 00
£ oo
«I ° P hJ
Oh’~
h4
O
Table 1.9-1
hi
hi hi hi
hi
hi
p hi
P
I cn I
i
3
Oh
P hi p ov p hi
-*->
3 a ''S’"
a"
o W
8 Hi sZIJ 8H oo 8Hj 8H
3 Ai
II II II II
a>
,,—^ /—S /—^
g
HH
hi
c
hi
si
3 hi hi
00
hi K
■J*
hi
‘-o p° P°
3 p p P
Dh
C
vs/
«W 8W! 8W’
r4i -se
i-H
S»/ *
Sec. 1.9 Superposition Summations for Discrete Systems 39
in this chapter, and the delta response will be written d(nT). For non-
anticipatory systems, d{nT) = 0 for n < 0. By superposition, the response
of a discrete system to the input of Eq. 1.8-2 is given by Eq. 1.9-1 in
Table 1.9-1. Equation 1.9-2 is obtained by replacing the dummy variable
k by n — k. Note that these equations define the output only at discrete
instants of time. For nonanticipatory systems with v(t) = 0 for t < 0,
the limits of summation become 0 and n for both equations, although it
is never incorrect to use the wider limits of — oo and + oo.
Consider a continuous system whose input is given by Eq. 1.8-4. The
output is the continuous function of time given by Eq. 1.9-3 or 1.9-4.
In the simplifications listed for nonanticipatory systems and for inputs
that are zero for t < 0, the statement “Upper limit is kT = is literally
correct only when t is an integral multiple of T. The summation is only
over discrete values of k. At time t = (n + y)T, where n is an integer
and 0 < y < 1, the statement above should be interpreted as k = n.
Equations 1.9-3 and 1.9-4 give the exact output of a continuous system
to a train of weighted impulses. When such an input is used to approxi¬
mate the output of a physical sampler, however, these equations likewise
involve an approximation. The degree of approximation depends upon
the width of the pulses from the sampler compared to the time constants
of the system.
If the output of the continuous system is desired only when t is an
integral multiple of T, Eqs. 1.9-3 and 1.9-4 reduce to Eqs. 1.9-5 and 1.9-6,
respectively. The latter equations would be used if, for example, the
output of the system were to be followed by another sampling switch.
For nonanticipatory systems with v(t) = 0 for t < 0, the limits of sum¬
mation again become 0 and n. The reader has no doubt noticed the
similarities between Tables 1.6-1 and 1.9-1. In fact, Eqs. 1.9-3 through
1.9-6 can be treated as special cases of Eqs. 1.6-2 and 1.6-4.
Equations 1.9-1 and 1.9-2 are identical with Eqs. 1.9-5 and 1.9-6,
respectively, except for the use of different symbols. It is therefore
instructive to compare the discrete and continuous systems shown in
parts (a) and (b) of Fig. 1.9-1. The relationship of y(t) to ift) in part (b)
(a)
(b)
Fig. 1.9-1
40 Time-Domain Techniques
Thus
y(n) = 0 for n < 1
2/(2) = 2
2/(3) = 8
2/(4) = 22
The calculation of y(n) for large values of n can be greatly simplified if the expression
can be put in closed form. It can be shown, although not without some thought, that
It is shown in Example 2.12-6 that the delta response d{k) in the last
example describes a system characterized by the difference equation
y{k + 2) — 3 y(k + 1) + 2 y(k) = 2 v(k + 1) — 2 v(k)
One disadvantage of the superposition summations in Table 1.9-1 is that
it is difficult and often impossible to express the answer in closed form.
When the preceding problem is solved in Example 3.13-2 by the use of
the Z transform, the solution is easily obtained in closed form.
Example 1.9-2. If, in Fig. 1.9-16, T = 1,
(2( for t > 1
hit) =
(O for / < 0
and
v(t) = tU^{t)
then the output at the instants t = n is
y(n) = 2[2n — n — 1] for n = 0, 1, 2, ...
Sec. 1.9 Superposition Summations for Discrete Systems 41
(a)
v(t) v*(t)
-> q(t) q*(t)
h(t)
y(t)
h(t)
(b)
Fig. 1.9-2
Example 1.9-3. Figure 1.9-2# shows a sampling switch followed by two identical sys¬
tems in cascade, each having the impulse response
Assuming that the second system does not load down the first, and that the sampling
interval is T = 1, find the output y(t) when
v{t) = <r*U-x{t)
From Example 1.6-2, the impulse response of the two systems in cascade is t3e~ll6.
Using Eq. 1.9-3,
-t t
y{t) = > e k--- -I (/ - *)’ for t > 0
k=0 6 fc=o
Thus
y(t) = Ue_V6 for 0 < t < 1
y{t) = U3 + (t — l)3]e~76 for 1 < t < 2
Example 1.9-4. Repeat the previous example, with the added sampling switch shown
in Fig. 1.9-2b.
m
y(t) = 2 (m - k) (t — m)e~ (t — m)
m=0 L fc=0
Using this approach, y(t) is not easily evaluated in closed form. For present considera¬
tions, it is sufficient to note that
It should be observed that the results of the last two examples are not
identical, not even at the sampling instants of t = n. Expressions for
y(t) in closed form are easily found by means of the modified Z transform
in Example 3.14-1.
42 Time-Domain Techniques
Co <[>
+3 X
ctJ C
ex' _■?
W Z
g .SS °
•2 - V
n g**- £ «
1.§
—1 uo
.3
X J5 s-,
o X || o
s—
i_ N<0 U,<u -v
cl s £
Ui <D J-H . 1-4 a>
<D N <L> <U N D
S<D O oo ^ oo cl 2.
X £ g-1 00 £ 00
ONi
• r-H ON
£
J-H
<D o o
• t-H
O cl.2
D
00
N X X X
X
e •" S' w
.2 g 2 -*-* •*«* a3 CO
X 0) <3
3 -<-» r>.
O >-.
oo .X1 •I o a> £ * I1 l_o
X L> 1—1 u* £
X 00 X J_<L> NO s_ £"< ctf »-ha> Lt* u u
Oh - c <3 ■$■ oo S-1 00 00 ^ <D N
e ! g
Cl, oo
O Cl oo ^O —1oo
CL'~ CL •-' CL’~
bo *5 o J X D D X
* Z
Table 1.9-2
a
a,
3
o
8
8
8(Aj I
I 8H,|
II
*
References 43
Time-Varying Systems
With these definitions, the equations of Table 1.9-1 are replaced by those
of Table 1.9-2 for time-varying systems. The interpretation of h*(t,kT)
as a weighting function is still valid.
REFERENCES
Problems
dny dy dmv dv
a”dr+'" + J, + dF< +••■ +b1Jt +b0v
2a
1.5 Find the response of Fig. 1.4-8 when the current source has the waveform
shown in Fig. PI.5, and also when ix{t) = 3e~t for t > 0.
h(t)
1.6 Check the results of Example 1.6-1 by considering the voltage source to
be the sum of two singularity functions, and by finding the response to
each one.
Problems 45
1.7 A fixed linear system containing no initial stored energy has the input
and impulse response shown in Fig. PI.7. Find the response at t = 4
u(t) h(t)
^12(0 = + T) dr
J — 00
C* 00
^21(0 = f2(r)fl(t + T) dr
J — 00
(7 + 1) + y = (t + 1M0
Noting that the left-hand side of this equation is the derivative of (t + 1 )y,
find hit. A), h*(t, t), r(t, A), and r+(/, r). Find the response to v{t) =
dy y dv
— -(-- == — -)- v
dt t + 1 dt
1.13 In the cascade connection of Fig. 1.2-1, assume that the impulse responses
of Nx and N2 are denoted by hx{t. A) and h2(t, A), respectively. Derive
an expression for the impulse response of the cascaded combination.
46 Time-Domain Techniques
1.14 Assume that three linear components are placed in cascade. If the first
is an ideal integrator, the third an ideal differentiator, and the middle
one characterized by the impulse response /z*(/, r) = te~iT+t) for r > 0,
find the impulse and step responses of the combination.
1.15 In Fig. PI.15, MO = (t - Ve-'U^it) and MO =* e^fcos (tt/2)/ -
sin (tt/2)/]I/MO cos [(tt/2)/ + (tt/4)]£/_i(0- The sampling switch
has a period T = 1 second. Find 2/0*70 when i>(0 = C/MO-
v(t) ^ 9*(0
h-i(t)
Fig. PI.15
1.16 A sampling switch is added just before MO Tig. PI. 15. The switches
operate synchronously with a period T = 1 second, and MO = (1/2)*
and MO = 2 cos (tt/3)/1 for t > 0. Find y(nT) when i?(0 = C/_2(0» and
also when v(0 = sin (tt /3)1.
1.17 In Problem 1.15, find 2/(0 for 0 < t < 3.
1. 8 A certain feedback control system samples the input r>(0 at t = nT, where
= J, and n — 0, 1, 2, .... The response to v(t) = U_x(t) is
2/(0 = (e-yT/6)[l — e~nT], where t = (n + y)T, and 0 < y < 1. Find
and sketch 2/(0 for 0 < t <2 T when i?(0 — e~2t
2
Classical Techniques
2.1 INTRODUCTION
equation! is
in—1
dny y . . dy . , dm'v , dv
an + an-1 t + • ' • + ax — + a0y = b m-h ’ ‘ + b1— + b0v
dt7 dt n- 1 dt dtm dt
(2.2-1)
Since the input v(t) is presumably known, the right side of this equation
can be represented by F(t), which is often called the forcing function.
dy dn 1y dy
a + an-i ~ \ + • ’ ' + ai~ + aoV — F(t) (2.2-2)
dtn dt 1 dt
For linear systems, the a/s and b/s cannot be functions of v or y but may
be functions of t. For fixed linear systems, these coefficients must be
constants. The proof of these two statements is similar to Examples 1.2-3
and 1.2-4.
A system’s differential equation may be given, or it may have to be
found from a model of the system. In the latter case, a set of simultaneous
differential equations are written directly from the model. The next
section gives an example and discusses the solution of the simultaneous
equations to yield a single equation relating y{t) and v(t).
The operator p is often used to indicate differentiation and is defined
byj
PMO] = y MO] (2-2-3)
dt
If c1 and c2 are constants,
dm+nv
pm(pn v) = pm+nv =
dtm+n
(2.2-4)
pm(c ffil + C2v2) = c1pmv1 + c2pmv2
(p + c/){p + c2)u = [p2 + (c1 + c2)p + cxc2]v
where m and n are non-negative integers. In fact, in most respects the
operator p may be treated as an algebraic quantity. The most notable
exception is that it does not, in general, commute with functions.
p(tv) 5* t(pv)
p(l\V2) ^ VjffD2)
For varying linear systems, where the coefficients are functions of time,
A and B are time-varying operators. This is shown by writing
„ de 3 (1
c1T +
clt u +
1 \
/ 3_
C2
_ de2
Ci—
dt
+
1
—
Rx
£2
Fig. 2.3-1
50 Classical Techniques
and
* ~
3 1
de v de 2 (\ 1\ 1
I—+—U+-
c'Hi + *S\ + (Cx + C2) ——|- e2 dt
at \Ri R2J L
-c,p
dt
Differentiating the second equation term by term to remove the integral sign, and using
the operator p = djdt, these equations become, after assuming for simplicity that all
resistances, capacitances, and the inductance have values of 1 ohm, 1 farad, and 1
henry, respectively,
(p + 2)e3 — (p + l)e2 ex
~(p2 + p)e3 + {Ip2 + 2p + l)e2 = (p2)e1
Premultiply each term in the first equation by the operator p2 + p, and each term in the
second equation by p + 2, and then add the two equations. Since
The procedure used in the preceding example is valid for any two time-
invariant equations. If L represents an operator which is a function only
of /?, the equations can be written symbolically as
Premultiply the first equation by L21, and the second by Ln, and subtract.
Since L2lL11y1 =
Each of the last two equations has only one independent variable and can
be solved by the methods of Section 2.6.
Consider next the general case of n simultaneous equations. Symbol¬
ically,
LnVi + L12y2 + * * • + Llnyn — F^t)
+ F22y2 T- • • • T* L2nyn — F2{t)
As long as the Lij operators depend only upon p, the solution obtained
by Cramer’s rule can be shown to be valid.f
^(p)yi=I.Ski(p)Fk(t) (2.3-5)
k= 1
E-12 ... r
^-11 ^ln
C2i ... T
A(p) = T22 ^2 n (2.3-6)
... T
Lni Ln2 nn
Aki(p) ^ the operator found by evaluating the kith cofactor of this deter¬
minant, i.e., the determinant A(p) with the A:th row and zth column
removed, multiplied by (—l)fc+h
This section has thus far been restricted to time-invariant differential
equations. Obtaining a single differential equation relating the output
and input of a varying system is more difficult. Suppose that a system
is represented by the simultaneous equations
Similarly, Eq. 2.3-5 does not hold. While a single differential equation
relating the output and input can often be obtained, no simple formula
of the type of Eq. 2.3-5 exists which is generally applicable. Also, the
difficulty in determining a single differential equation often makes other
methods advisable. One of these methods is discussed in Chapter 5.
f Those in need of a review of determinants and Cramer’s rule should read Sections 4.3
and 4.6, respectively.
52 Classical Techniques
Any «th order linear differential equation can be written in the form of
Eq. 2.2-5, repeated below.
Because all the remarks of this section are valid for both fixed and varying
systems, tjae a/s can, in general, be functions of t. If the right side of the
last equation is identically zero,
where the K/s are arbitrary constants. The subscript H refers to the
solution of the homogeneous equation.
The last equation says that, once n independent solutions are known,
any other solution can be expressed as a linear combination of these n
solutions. For fixed systems, there is a general method of finding n
independent solutions to the homogeneous differential equation. For
varying systems there is, unfortunately, no such general method.
The most general (or “complete”) solution of the nonhomogeneous
Eq. 2.4-1 is
y = yH + yP (2.4-5)
Sec. 2.5 Solution of First Order Differential Equations 53
where yH is given in Eq. 2.4-4 and is the solution of the related homo¬
geneous equation. yP is any one solution, no matter how arrived at,
which satisfies Eq. 2.4-1, and is known as the particular or particular
integral solution. yH is called the complementary solution. Any method
of finding yP, including guesswork, is allowable. In fixed circuits with a
sinusoidal source, for example, a-c steady-state circuit theory might be
employed. The next section shows an explicit general method of finding
yP once yH is known.
As yP contains no arbitrary constants, y and yH both contain n such
constants. Their evaluation requires a knowledge of initial or boundary
conditions, and it is discussed in Section 2.6.
Any linear first order differential equation can be written in the form
of Eq. 2.5-1.
For convenience, it is assumed that the coefficient of dyjdt has been made
equal to unity. The coefficient a may in general be a function of t. Such
an equation may always be solved by the introduction of the integrating
factor eJa(t)dt. Both sides of Eq. 2.5-1 are multiplied by this factor.
The left side of the last equation is the time derivative of y€lait)df So
Although the integration on the right side of Eq. 2.5-3 may sometimes be
difficult, the method constitutes an explicit procedure for solving the
differential equation, for both the time-invariant and the time-varying
cases. It will give both the complementary and the particular solutions
in the answer.
In evaluating |'a(t) dt, it is not necessary to include an arbitrary con¬
stant of integration. The reader should convince himself that including
such a constant will not increase the generality of the final solution.
54 Classical Techniques
4
dt
= tc*'*
y — — 1 + ce*2/2
If the rf s are all different, it turns out that W(t) in Eq. 2.4-3 does not
vanish, so that the n individual solutions are independent. If r1 = r2,
then yx = erit and y2 = terii are independent solutions. If a root, say rl9
is repeated k times, so that rx — r2 = • • • = rk, then the most general
solution is
yH = Kxer i* + K2t€ri* + • • • +
Thus finding yH involves only solving for the roots of an nth order equa¬
tion. In the event that some of the roots are complex, however, the
solution should be written in another form.
Since the coefficients in Eq. 2.6-2 are real, any complex roots must
occur in complex conjugate pairs. If one root is r1 = a + jfi, where
a and ft are real, another root must be r2 = a — jf. Then
whose solution is
V = Vn + Vp (2.6-6)
The complementary solution yH is found by replacing F(t) by zero and
solving the resulting homogeneous equation as before. There are two
standard methods of solving for the particular integral solution yP, the
56 Classical Techniques
g + 3^ + 2 y=l+c-‘
dt2 dt
for which yH = + A2e_2<, is not identically satisfied by yP =
A + Be~\ regardless of the choice of A and B. It is reasonable to expect,
however, that the term resulting from exciting the system at a natural
mode would decay more slowly with time than would otherwise be the
case. It is thus logical to try as a solution
yP = A + Bter1
d2y dy
— + 2 — + y = tc-(
dt2 dt
yP = At3e~l + Bt2e~f
Sec. 2.6 Solution of Differential Equations with Constant Coefficients 57
Note that, although repeated differentiation of At3e_i also leads to the terms C/e_< and
De~\ these terms are not included in the assumed yP. The reason is that these terms
are solutions of the related homogeneous differential equation and would therefore
vanish when they are substituted in the left side of the original differential equation.
Substituting yP as determined above,
has only the single term yH = Kyx. The particular integral solution is
assumed to have the form
yP = uy1 (2.6-9)
where all three symbols in Eq. 2.6-9 are functions of t. To find u, Eq.
2.6-9 is substituted into Eq. 2.6-7, giving
dy e 31
— + 3y = —
dt t
du g( e~3t 1
dt t t
u = In t and yP = Ke~3t In t
The complete solution is
y = Ke~3t + e~3t In t
Rearranging,
a2(Miih + w2</2) + ul(a2ij1 + a1yl + aay^
+ u2{a2y2 + a1y2 + a0y2) = F(t)
Since y2 and y2 both satisfy Eq. 2.6-12,
m
«i*/i + u2y2 = (2.6-15)
Q2
To obtain explicit formulas for ux and u2, Eqs. 2.6-14 and 2.6-15 are solved
simultaneously. There results
Wi = -*/2r(0
u« =
y,m (2.6-16)
~ ViVi) a2(2/i*/2 - ViVs)
It is worth noting that, since yx and y2 are two independent solutions of
Eq. 2.6-12, Eq. 2.4-3 implies that yxy2 — y1y2 ^ 0. Since the denominator
of Eq. 2.6-16 is not zero, u1 and u2 can always be explicitly found.
Example 2.6-3. Find the complete solution to
d2y 1
-T7 + y = -
Vn ~ Ki cos t + K2 sin t
Thus
yx = cos t and y2 = sin t
In Eq. 2.6-16,
2/i2/2 - ViVi = cos21 + sin2 t = 1
sin t cos t
-, «2 =-
t t
The complete solution is
sin t cos t
y = Kx cos t + K2 sin t — (cos t) dt + (sin t) dt
The integrals appearing in the answer cannot be expressed in terms of a finite number of
elementary functions. The integrals can be expressed by infinite series, however, and
happen to be well-tabulated functions—the “sine integral” and “cosine integral”
functions that have many applications.
Consider finally the n\h order differential equation shown in Eq. 2.6-5.
The complementary solution has the form
where the u - s are functions of /. The derivatives of the u/s are found by
solving simultaneously the following n equations.
n—1 Fit)
UiVi + umI 1 + • • • + unynn 1 =
an
where dots and superscripts denote derivatives with respect to t. The first
w — 1 of these conditions are arbitrarily selected to put the result in a
tractable form. The last of the equations is found by substituting the
assumed yP into Eq. 2.6-5 while making use of the first n — 1 equations.
The set of equations above can be solved in terms of determinants by
Cramer’s rule.
WJt)F(t)
ui for i — 1, 2, . . . , n (2.6-18)
anW(t)
where W is the determinant
Vi y2 • • • Vn
Vi 2/2 • • • Vn
mt) = (2.6-19)
. ,71—1 . .n—1 „ .n—1
Vi y2 Vn
and where Wni(t) is the mth cofactor. By Eq. 2.4-3, W(t) is the Wronskian,
which does not vanish if yx through yn are independent solutions of the
homogeneous equation.
Example 2.6-4. Find the complete solution to
d3y d2y
+ 3 + y =
dt3 dt2
The homogeneous solution is
yH = K.e-t + KtU~* + K^
1 t t2
W(t) = e~3< -1 (—/ + 1) (->2 + 20 2e~3t
1 (t - 2) (/2 - At + 2)
ux = (|e3<)(/26-29(e-9 = bj2
u 2 = ae39(-27e-2t)(e-9 = -t
*3 = (*€80(€-a*)(e-*) = *
«i = h3
U2 = —¥2
«3 = y
yP = (i?3)C-4) + (—*/*)(/€-*) + t3e -t
Sec. 2.6 Solution of Differential Equations with Constant Coefficients 61
If v(t) = est, the particular solution has the form yP{t) = H(s)est, where
H is not a function of t. Substituting these expressions into the differential
equation gives
A(/?)[//(T)esi] = 5(/?)[esi]
In the particular solution for v{t) = est, all currents and voltages have the
form e(t) = E(s)est and i(t) = I(s)est. The three passive circuit elements
are shown in Table 2.6-1. When the preceding expressions for e and i are
Table 2.6-1
R e = Ri ' R
—wv—
di
e =L — sL
dt
1
c 1 = c dt— sC
—ii—
inserted into the defining equation for the circuit element, the ratio of the
voltage to the current can be found. This ratio, called the impedance
and denoted by Z(s), is independent of t. Because of this fact, the relation¬
ship between yP(t) and v(t) can be found by the same basic rules as apply
to d-c circuits.
Example 2.6-5. For the circuit of Fig. 2.6-1, find the particular solution eP(t) when
the input current is i(t) — est.
(R + sLXl/sC) sL + R
E(s) = -I(s) =- I(s)
R + sL + 1 /sC s2LC + sRC + 1
Thus
sL T R
H(s) =
s2LC + sRC + 1
sL T R
eP(t) = —-est
P s2LC + sRC + 1
Fig. 2.6-1
Sec. 2.6 Solution of Differential Equations with Constant Coefficients 63
If the complex quantity H(jco) is expressed in polar form by \H(jco)\ eje, the
response can be written
yP = \H(j<o)\ cos (cot + 6)
H(jco) is often called the frequency spectrum of the system and is basic
to a-c steady-state circuit theory.
For negative values of t there is no stored energy associated with L and C. Find the
output voltage eft).
64 Classical Techniques
The circuit can be described mathematically by writing two node equations. Applying
KirchhofTs current law at nodes 2 and 3,
de2 1 \ 1
C17 + Rez I ~Re°=
1 / 1 1
e-> dt\ =0
~Re2+\Rea + L
The second equation can be differentiated term by term to remove the integral sign.
Solving the first equation algebraically for e3 and substituting the result into the second
equation gives
d2e2 Rde2 1 _ ldex R
~dF + L~dt + LC62 ~ ~C \~dt +
As expected, this is the same differential equation as the one found for Fig. 2.6-1 by a
different method. Inserting the given numerical values and expressing / in microseconds
rather than seconds, there results
The particular integral solution can be found by examining the differential equation,
or by using d-c steady-state theory.
14,400
02)p = -18
800
The complete solution is
Also
= e-18t[(21.8i5r2 - 187Q cos 21.St - (21.8Kx + 18AT2) sin 21.8?] for t > 0
at
Kx and K2 can be evaluated from the last two equations if e2 and de2/dt are known at
t = 0 + . From the problem statement, the voltage e2 and the current through L were
zero for negative values of t. These values cannot change instantaneously unless there
is an impulse of current into C or an impulse of voltage across L, so they must remain
zero at t = 0 + . To find the value of de2/dt at t = 0 + , note that, since the inductor
current remains zero, the current through C must be ic = i = —gm = —0.01. Since
4 = C (defdt),
Example 2.6-7. Repeat the previous example with eft) = (cos 4 x \Wt)U_ft).
The form of the complementary solution is not changed, and the particular solution
can be found from the system function
jcoL -f- R
Hi j(o) = -4-
7 1 - co2RC + JcoRC
H(jco) = (—13.1)e-m°
eft) = -13.1 cos (40/ - 71°) + ^18t[Kx cos 21.8/ + K2 sin 21.8/]
The arbitrary constants can again be evaluated from e2(0+) and (de2/dt)(0+).
2/(0) = f
dt
(0)
d'-'y
dt"-1
(0) = 0 (2.6-24)
66 Classical Techniques
where y±(t) and y2(t) are two solutions of the homogeneous equation, and
where W(t) is the Wronskian
The previous comments about the upper and lower limits of integration
still apply. Note that y(0 ) = 0, and
+ [j/i(0“i(0 + 2/2(0“2(0]
KM F(z) dz (2.6-27)
= 1
0 an(z)W(z)
where W(t) and Wni{t) are defined in Eq. 2.6-19. Since they are based on
the variation of parameters method, Eqs. 2.6-24 through 2.6-27 are valid
for varying as well as for fixed systems. There is an important application
in Section 2.8, where these equations are used to find the impulse response
of a time-varying system.
Sec. 2.6 Solution of Differential Equations with Constant Coefficients 67
The form of the complementary solution depends only upon the system
and not upon the input. The characteristic equation depends only upon
the parameters of the system, and the roots of the characteristic equation
determine the kind of terms appearing in the complementary solution.
In the event that there is no external source (e.g., the system may be excited
only by some initial stored energy within it), the complementary solution
becomes the complete solution. Thus the complementary solution
represents the natural behavior of the system, when it is left unexcited.
For this reason, the complementary solution is also called the free or
unforced response.
If the free response of a system increases without limit as t approaches
infinity, the system is said to be unstable. This is the case if the character¬
istic equation has a root with a positive real part, since the complementary
solution then contains a term which increases exponentially with t. Roots
with negative real parts, on the other hand, lead to terms that become
zero as t approaches infinity. Purely imaginary roots, if they are simple,
lead to sinusoidal terms of constant amplitude in the complementary
solution. This case of constant amplitude oscillation in the complemen¬
tary solution, which is characteristic, for example, of LC circuits, is usually
considered a stable and not an unstable response. Repeated imaginary
roots lead to terms of the form tn cos (ojt + <j>), which is an unstable
response. If the roots of the characteristic equation are plotted in a com¬
plex plane, the following statement can be made. For a stable system,
none of the roots can lie in the right half-plane, and any roots on the
imaginary axis must be simple. If all the roots of the characteristic
equation lie in the left half-plane, the complementary solution approaches
zero as t approaches infinity and is identical with the transient response
of the system.
The magnitudes of the terms in the complementary solution, i.e., the
arbitrary constants, depend upon two things, one of which is the input.
The other is the past history of the system before the input was applied.
This history can be completely summarized by a knowledge of the energy
stored within the system at the time the input is applied.
The form of the particular solution is dictated by the forcing function,
as can be clearly seen from the method of undetermined coefficients.
The only time the system has any influence upon this form is when a term
in the forcing function duplicates a term in yH. In this event, the system
is being excited in one of its natural modes, a phenomenon called reson¬
ance.
68 Classical Techniques
Since the form of the particular solution depends upon the input, it is
also called the forced response. If all the roots of the characteristic
equation lie in the left half-plane, the forced solution is identical with the
steady-state solution. The magnitudes of the terms in the forced solution
depend upon both the input and the system parameters.
It is customary to think of the forced component of the solution as
being immediately established by the application of the input. The free
component, which is the complementary solution, adjusts itself through
the proper evaluation of the arbitrary constants to provide for the proper
transition from the unexcited system to a system under the dominance
of the input. Some like to think of the system as initially resisting
the wishes of the input, by means of the complementary solution. The
magnitudes of the arbitrary constants depend upon how greatly the
character of the input differs from the natural behavior of the system.
/ x dny , , ,. dy , , N
an(0 ~rp + ’ ' ‘ + a1(0 - b a0(t)y
dt dt
For first order equations, the method of Section 2.5 always yields a
solution. For higher order equations, however, an explicit solution
cannot be found, in general.
Consider Eq. 2.7-1 with F(t) replaced by zero, and examine the method
of Section 2.6 for finding yH. Solutions of the form y — ert, where r is a
constant, were assumed, resulting in the equation
anrn + • * • + ay + a0 = 0
But, since the a's are now functions of time, the roots of this equation
are functions of time, violating the assumption just made. In fact, there
is no general method of finding yn in terms of elementary functions.
See. 2.7 Solution of Differential Equations with Time-Varying Coefficients 69
cl2y dy
t2 77 + t—+y = °
dt2 dt
Let e3 = t and e2 dz = dt. Then
dy dy
dt dz
dt2 dz dz2-
The new equation becomes
d2y
- +y = o
dz
the solution of which is
The first of these two equations may be solved for v as a function of t, and the second
then solved for y, using the method of Section 2.5.
(Od/O 1
— and u2 = —tin/
t\-2/t)
1
Vp = “iVi + uyjz =
4~t
y = Ky + — (K, - 4i 1 In /)
t
t See Section 5.9.
Sec. 2.8 Obtaining the Impulse Response from the Differential Equation 71
The identification of the complementary and particular solutions with the free and
forced response is still valid for time-varying differential equations.
Since the system is at rest for t < 0, y = 0 for t < 0. Because the forcing
function U0(t) is zero except at t = 0, the solution for all t > 0 consists
only of the complementary solution
V = K1y1 + K2y2 + • • • + Knyn (2.8-2)
To evaluate the n arbitrary constants, n initial conditions are needed.
For a fixed system, where the a/s are constants, these can be found directly
from the coefficients. The nth derivative of the solution, but none of the
lower order derivatives, contains an impulse at t = 0. This is the only
way Eq. 2.8-1 can be satisfied at t = 0. If one of the lower order derivatives
were to contain an impulse, then dnyjdtn would contain a singularity of
higher order. Since an{dnyjdtn) does contain a unit impulse at t = 0,
dn~Yy\dtn~Y must jump from 0 to 1 \an at t = 0, and all lower order
derivatives must be continuous at the origin. The n initial conditions must
be
1
2/(0 + ) = 2/(0 + ) = • • * = 2/ (0 + ) = 0, 2/n-1(0+) = - (2.8-3)
a‘n
„
d2y \
—— = cot/^) + CiU0(t) + [2c2 + * • • ] C/Li(r)
dt2
From the original differential equation, the complementary solution for t > 0 is
y = Kye-' + K2e~2t
The reader familiar with the Laplace transform may wish to compare this example with
Example 3.5-2.
Note the similarity of this equation and Eqs. 2.6-25 through 2.6-27. Since
yft) is not a function of 2, and F(z) and W(z) do not depend upon i, Eq.
2.6-27 may be rewritten as
n
1
-a„(z)W(z)
2 yi(‘Wni(z) dz (2.8-7)
In general, F(t) 9^ v(t), and Green’s function and the impulse response
are not identical. Let the only restriction on Eq. 2.8-5 now be m < n.
Fit) = B(p, t)v(t)
= [bm(Opm + • • • + W{t)p + b(1(t)]v(t) (2.8-12)
where p = djdt. Equation 2.8-9 can be used to find the impulse response,
once Green’s function is known. If v(t) = U0(t — 2), then y(t) = h(t, 2).
where now p = djdz. Equation 2.8-13 is used only to find the impulse
response for 0 < 2 < t, and the integrand is zero except at z = 2. The
limits on the integral may therefore be changed to —00 and +00.
This equation is not nearly so formidable as it looks. A typical term in
B(p, z)U0(z - 2) is
Thus yx{t) = 1 It and y2(t) = 1 It2. To determine Green’s function, note that
1 1
z z2 ]
W(z) =
-1 -2
z2 z3
1 1
7 T2 z — t
2 »<(ok, = (-i)
i= 1 i i t2z2
z z2
,l\ (z — t \ t — z
6cr <'•*>“ I?)(-*‘>(-Fv t2
t - ?.
h(t, 7) = —-— for t > X
A = E - 1 (2.9-4)
Ay(k) is called the first difference of the function y(k). Higher order
differences are defined as
or, in general, \
A[cy(k)] = c Ay(k)
A m[y(k) + z(k)] = A my(k) + A mz(k) (2.9-6)
Am Any(k) = An Am?/(^) = Aw+n?/(A:)
m = hmfA±iCim (2.9-7)
dt t~>o T
m) =/<< + t)
(2.9-8)
A/(0 =f(t + T) -fit)
Then
(2.9-9)
dt t-+ o T
Sec. 2.9 Difference and Antidifference Operators 77
and, in general,
dmf (0 _ lim AW/(0 (2.9-10)
dtm T-+ 0 Tm
It is not surprising, therefore, that there are differencing formulas similar
to but not identical with the common differentiation formulas. For
example,
\[y{k)z(k)] = y(k + 1) Az(/c) + z(k) Ay(k)
P_1[/(01 = J/( 0 dt + c
= I /(A) dl + K (2.9-14)
Jt0
where c and K are constants of integration. The lower limit t0 is arbitrary
and forms part of the constant of integration. Specifically,
c = K - f(t) dt
_ %) _n=fo
The quantity
2/(0 = p-\m\ (2.9-15)
MO =/(0 (2.9-16)
Equivalently,
PP~V(0 =/( 0 (2.9-17)
where the summation is still with respect to the dummy variable n. The
lower limit in Eq. 2.9-22 is unspecified because any fixed number of the
terms/(0),/(l),/(2), ... in Eq. 2.9-21 can be combined with the constant
of summation, K, to form the new constant c. The lower limit may be
selected in whatever manner appears to be the most convenient. The
arbitrary lower limit in Eq. 2.9-22 is the analog of the arbitrary value of
t0 in Eq. 2.9-14.
The reader may recall that the exact evaluation of integrals may be
tricky or difficult, and in some cases downright impossible. Since there is
no conceptual difficulty in summing a finite number of terms, he may be
tempted to conclude that a similar problem does not exist for the operator
A-1. A brute force calculation of A~lf(k) from Eq. 2.9-22 over a wide
range of values of k is often, however, unnecessarily tedious. As k
increases without limit, so does the number of terms in the summation.
Whenever possible, A~xf(k) should be expressed in closed form.
The summation of a finite series is one of the topics usually included in
books on finite differences. Among the available techniques are short
tables of summation formulas, summation by parts (analogous to but
not identical with integration by parts), the use of Bernoulli polynomials,
and the use of partial fractions. For example, the telescopic series
n=2 n(n — 1)
Sec. 2.10 Representing Discrete Systems by Difference Equations 79
1 = _L_1
n(n — 1) n — 1 n
The result is
k i i
2 -= 1 - - (2.9-23)
n=2 n{n — 1) k
n—k
The summation of some of the very simple function (e.g., 2 l/«) cannot be
expressed in closed form in terms of elementary functions. In some of
these cases, an approximate summation can be obtained.
It is useful to note that, when factorial polynomials are used, Eq. 2.9-13
gives
A = —-— (/c)(m+1> + K
m + 1
or
A-\k{k - 1) • • • (k - m + 1)]
n=k—1
The general form of a difference equation relating the output y(k) to the
input v(k) of a discrete system is
or, equivalently,
Either of the two equations may be easily obtained from the remaining
one by Eqs. 2.9-1 through 2.9-5. The second is a closer analog to the
differential equation 2.2-1, but the first is easier to use and is the one more
commonly given. For a linear system, the a/ s, b/s, c/s, and d/s cannot
be functions of y or v, but they may be functions of k. For fixed linear
systems, these coefficients must be constants.
Using the shifting operator E, Eq. 2.10-1 may be rewritten as
Since the input v(k) is presumably known, the right side of Eq. 2.10-1
is represented by the known forcing function E(k). For fixed linear
systems, where the coefficients are constants, the last equation is written
symbolically as
A(E)y(k) = B(E)u(k) = F(k) (2.10-4)
where k, m, and n are integers. By this equation, the value of the output at
the instant t — (k -f n)T is expressed in terms of the past outputs from
t = kT to (k + n — 1 )T, and in terms of the inputs from t = kT to
(k + m)T. For nonanticipatory systems, m < n. If the time scale is
adjusted so that T = 1, this equation is identical with Eq. 2.10-1.
Difference equations also arise in ways other than those mentioned
above. The right side of Eqs. 2.10-3 and 2.10-4 need not always be a
direct function of the input, and the discrete variable k may be an index
of position instead of an index of time. The following two examples!
show how difference equations result whenever there is a repetition at
equal intervals of position or at equal intervals of time.
Fig. 2.10-1
Fig. 2.10-2
While the solution could be found by calculating the response for each cycle in
succession, this would be a very tedious process. The steady-state solution could be
obtained by use of the Fourier series, but the answer would be an infinite series instead
of being in closed form. Examine, therefore, the £th cycle, and for convenience move
the time origin to the start of the Ath cycle. The differential equations describing the
circuit are
di1
-b i1 = 1 for 0 < t < 1
dt
dio
-b /2 = 0 for 1 < t < 2
dt
where the subscripts 1 and 2 denote the first and second halves of the cycle, respectively.
Since the current is a function of the cycle under consideration, it is again a function of
82 Classical Techniques
the discrete variable k as well as the continuous variable /. The solution is carried out
in Example 2.12-3. A first order difference equation must be solved, along with the
above differential equations.
t This is the basis for one method of obtaining numerical solutions of differential
equations.
Sec. 2.12 Solution of Difference Equations with Constant Coefficients 83
y± y2 • • * yn
Ey1 Ey2 • • • Eyn
W(k) = * ot (2.11-5)
y = yH + yp
where yH is given in Eq. 2.11-6, and where yP is any one solution satisfying
Eq. 2.11-1. The components yH and yP are called the complementary
and particular solutions, respectively. Since yP contains no arbitrary
constants, y contains n such constants, which must be evaluated by initial
or boundary conditions, as in Examples 2.12-1 and 2.12-4.
There must be two independent solutions to this second order difference equation. For
the assumed solution i{k) = erk,
R1 + R2
cosh r
R*
Since cosh ( — 0) = cosh (0), the two allowable values of r are given by r = ±0 where
+ R2
(2.12-4)
cosh 0 =
R*
The general solution is
i(k) = Cxckd + Co€~k6
The two boundary conditions needed to evaluate the arbitrary constants Cx and C2
are found by examining the first and last meshes in Fig. 2.10-1.
Co = -Cxe27n0
and
i(k) = CA^9 - e2mee~ke] = —2Cxcm9 sinh (m - k)0
for k = 0, 1, 2, . . . , m, where 0 is given by Eq. 2.12-4. Example 3.7-3 gives the solution
for i as a function of t and k when the resistance R2 in Fig. 2.10-1 is replaced by a
capacitance C.
While the form of the assumed solution given in Eq. 2.12-2 happens to
be the most convenient one for Example 2.12-1, it is not the form that is
usually assumed. Recall from the example that the allowable values of r
were the roots of a polynomial in er. If p = €r, the assumed solution
becomes
y(k) = pk (2.12-7)
If the ftfs are distinct, it can be shown that the individual solutions
yx = Pf, y2 = p2k, . . . , yn = pk do satisfy Eq. 2.11-5 and are therefore
independent.10 If a root, for example, is repeated m times, then the
most general solution is
Vp = [Ak3 + M2](-l)*
Then
EyP = —[A{k + l)3 + B(k + l)2](-l)fc
E2yP = [A{k + 2)3 + B(k + 2)2](— l)fc
Sec. 2.12 Solution of Difference Equations with Constant Coefficients 87
Substituting these expressions into the left side of the difference equation yields
where the subscripts 1 and 2 denote the first and second halves of the cycle, respectively.
The solutions of these differential equations are
The factors Cx and C2 are constants with respect to t but may be functions of k, the
cycle under consideration. Since the current cannot change instantaneously,
4(0, 1) = 0
Inserting each of these three conditions into Eq. 2.12-15 gives, respectively,
C2(k) = CM + e (2.12-16)
1 + Cffc + 1) = C2(k)e-2 (2.12-17)
Cx(l) = -1 (2.12-18)
Equations 2.12-16 and 2.12-17 yield the first order difference equation
Cffc) = Ce —
1 + e-1
88 Classical Techniques
By Eq. 2.12-18,
1 + e
Finally,
e + e2(1-fc)
Cx(k) = -
1 + €
and
€ + e2(1-*> ,
i1(7, k) = 1--—- e~* for 0 < t < 1
1 + e
e2 _ e2(l-fc)
e
4(0 = 1 - for 0 < t < 1
1 + €
(2.12-19)
= F(k) (2.12-20)
The complementary solution has only the single term yH = Cy^k). The
particular solution is assumed to have the form
equation. Therefore
aiVi(k + 1) Afx{k) = F{k)
whose solution, by Eq. 2.9-22, is
fi{k) = A
-i F(k) f F(n - 1) (2.12-22)
Lfli2/i(fe + l)-l «i2/i(n)
Example 2.12-4. Find the complete solution to
(-3)1
(£ + 3 M*) =
/:(A: +1)
The complementary solution is
, __
yp = jU(/c)(-3)k
By Eq. 2.12-22,
(_3 1
(-3)" ,4 -3(« - 1 )n
By Eq. 2.9-23,
1
ffk) = -i
1 1 1
y(k) = C-T — (— 3)fc = Ci + — (-3)fc
3 3 k_ 33K_
k_
Vh = Ctfxik) + C2y2(k)
The particular solution is assumed to be
One of the two conditions needed to evaluate and fx2 is that Eq. 2.12-24
must satisfy Eq. 2.12-23. By analogy to Eq. 2.6-14, the second condition
is arbitrarily chosen to be
Vi(k + 1) Anfk) 4- y<i(k + 1) Aju2(k) = 0 (2.12-25)
Since
ix iff + 1) = fxfk) + A fxfk)
and
E2Vp = Vi(k + 2 )[yx(k) + A^*)] + y2(k + 2)[y2(k) + ky2(k)}
1
Substituting these expressions into Eq. 2.12-23 and rearranging gives
Since yx(k) and y2(k) are solutions of the related homogeneous equation,
Similarly,
jl = ]c 71 — Jc
Mk) = -2 (-!)“(» - 1X-1)”-1 =2(n - 1) = ik(k - 1)
Then
yP = -Kk3 - k)(-l)k + K£2 - k)k(-iy
= [ik* - W2 + ^](-i)fc
The complete solution is
Consider finally the nth order difference equation, Eq. 2.12-12, whose
complementary solution has the form
(2.12-31)
Solving Eqs. 2.12-30 and 2.12-31 in terms of determinants by Cramer’s rule,
Wni(k + 1 )F(k)
(2.12-32)
a„W(k + 1)
or
r=k
Wni(r)F(r - 1)
ft =1 anW(r)
(2.12-33)
and where Wni(k) is the mth cofactor. By Eq. 2.11-5, W(k) does not
vanish if yx{k) through yn(k) are independent solutions of the homogeneous
equation.
Although the summation involved may be difficult to express in closed
form, Eq. 2.12-33 gives an explicit solution for yP once yH is known.
Unlike the method of undetermined coefficients, the variation of param¬
eters method is applicable for all forcing functions and can be extended
to the solution of difference equations whose coefficients are functions
of k.
\
The form of the complementary solution depends only upon the param¬
eters of the system, specifically upon the roots of the characteristic
equation
A(0) = 0
If the roots f$2, . . . , f$n are distinct, the complementary solution
Since the form of the particular solution depends upon the input, it is
also called the forced response. If all the roots of the characteristic
equation are inside the unit circle, then the complementary and particular
solutions are identical with the transient and steady-state solutions,
respectively.
Some insight into the relationship between the classical and transform
solutions of difference equations is gained by examining the particular
solution of Eq. 2.12-35, which is written out below, when v(k) = zk.
Assuming that yP(k) = H(z)zk, where H(z) is not a function of k and noting
that EmyP(k) = H(z)zmzk, there results
H{z)A{z)zk — B(z)zk
where A(z) is the function formed from A(E) by replacing the operator
E by the parameter 2. Thus
bmzm + ’ • • + bxz + b0
(2.12-37)
anzn + • • • + axz + a0
H(z) is the system function for a discrete system and can be written down
by inspection of the difference equation. It plays an important role in
the application of the Z transform. Note also that
'yp(k)
H(z) (2.12-38)
- v(k). v(k)=z
The delta response, defined in Section 1.9 and denoted by d(k), is the
response of a discrete system initially at rest to
T for k — 0
(2.12-39)
0 for k Z 0
The right side of this equation is zero for k > 1, sq the delta response is the comple¬
mentary solution
y(k) = Ci + C2(2)k for k > 1
To evaluate the two arbitrary constants, the values of y( 1) and y(2) are needed.
Substituting k = —2 in the difference equation gives
y{ 0) — 0 + 0 = 0 — 0
y( 1) — 0 + 0 = 2 — 0
Then y( 1) = 2. For k = 0,
2/(2) - 3(2) + 0 = 0-2
Thus 2/(2) = 4. Using the last two results to evaluate Cx and C2,
2 = Ci + 2C2
4 = Cj + 4C2
= «/(0) IT a(n)
n=0
The lower limit of the indicated product is arbitrary, because any fixed
number of the factors a(0), <?(1), a(2), . . . may be combined with the
arbitrary constant C.
The solution to a homogeneous equation of order greater than one
cannot in general be found in terms of elementary functions, since the
procedure based on Eqs. 2.12-7 and 2.12-8 is invalid when the coefficients
are functions of k. It is sometimes helpful to know that, if all but one
of the independent solutions are known, the remaining one can then be
found.11
As with differential equations, there are a number of special cases
which do have an explicit solution. If the equation can be put in the
form
may be written as
(kE + 1)(£ + k)y(k) = 0
(kE + 1 )z(k) = 0
(E + k)y(k) = z(k)
[(* + 1 )E - k]y(k) = k + 1
The particular solution is assumed to have the form yP — /i(k)/k. By Eq. 2.12-22,
where the last expression follows from the formula for an arithmetical progression or
from Eq. 2.9-24. The complete solution is
C k + 1
y(k) = - +
k 2
REFERENCES
1. A. Ralston and H. S. Wilf, Numerical Methods for Digital Computers, John Wiley
and Sons, New York, 1960, Part III.
2. E. L. Ince, Ordinary Differential Equations, Dover Publications, New York, 1956,
Section 5.22.
3. Ibid., Section 7.4.
4. Ibid., Section 15.7.
Problems 97
Problems
Fig. P2.3
98 Classical Techniques
2.4 Find the most general solution of the following homogenous differential
equation by noting that one solution is y^t) = t.
(\t3p2 + tp - \)y = 0
2.5 Solve Example 1.6-1, using only the classical solution of differential
equations.
2.6 The circuit in Fig. P2.6 is originally operating in the steady state with the
switch K open. If the switch closes at t = 0, find expressions for the
currents for t > 0. \
112 2h
VA-rWfx
+
4 volts -=■
2a 112
< •— W\r AAA--
+
2 volts —
Fig. P2.6
2.7 Find the impulse response of the systems described by the following
differential equations.
(c) (p + t)y = v
(d) (tp2 + 2p)y = v
2.8 Solve Problem 2.2d by using the superposition integrals of Chapter 1 and
the result of Problem 2.1d.
2.9 Find an expression for h(t, X) if a system is described by (ap + \)y =
(ap + \)v, where a is a constant, but where a may be a function of t.
2.10 Prove Eqs. 2.9-6, 2.9-11, and 2.9-13.
2.11 Sum the series
i —-—
"(n + 1)(« + 2)
2.12 Starting with Eq. 2.9-11 a, derive the formula for summation by parts:
k k
2 u{n) A ( ) = [u(n)v(n)]^+1 —
77 77 2 v(n + 1) &u(n)
Problems 99
2.13 Find the most general solution of the following difference equations.
Which of these equations represent stable systems? Rewrite these equa¬
tions in terms of the difference operator A.
2.14 Solve the following difference equations, and comment on the stability
of the systems they represent.
(a) (E - k)2y(k) = k!
0b) [E2 + (2k + \)E + k2]y(k) = k\
Transform Techniques
3.1 INTRODUCTION
f2(t)
fM
An infinite series, known as the Fourier series, may be written for any
single-valued periodic function which has in any one period only a finite
number of maxima, minima, and discontinuities. The area underneath
any one cycle of the curve is required to be finite. The Fourier series for a
function /(f) with a period T is
00
where co0 = lirjT is called the angular frequency of the first harmonic,
and where 1 ot/2
a0 — /(0 dt
T J-T/2
2 0TI2
an /(f) cos nco0t dt (n 0) (3.2-2)
T J-T/2
2 fT/2
bn = - /(f) sin nco0t dt
T J-T/2
The series converges to /(f) for values of t at which the function is
continuous. At a finite discontinuity, the series converges to the average
of the values of/(f) on either side of the discontinuity. The series con¬
verges uniformly for all f, if/(f) satisfies the restrictions of the previous
paragraph and in addition remains finite. If only a finite number of terms
are taken, the corresponding finite Fourier series approximates /(f) with
the least mean square error.
Since two trigonometric terms of the same frequency may be combined
into a single term with a phase angle, Eq. 3.2-1 may be rewritten as
An and 0n represent the magnitude and phase angle (with respect to a pure
cosine wave) of the nth harmonic.
m= 2 n=— oo
c,**-** (3.2-6)
where
1 fTI 2
C« = f dt (3.2-7)
1 J-T/2
C, = ~ (n * 0)
eo (0
104 Transform Techniques
For the calculation of the steady-state response, the square wave of input voltage
may be assumed to exist for both positive and negative values of time. Then the
frequency spectrum of the input, found from Eq. 3.2-7 with co0 = n, is
1 _ £—jmr
Ijnrr
cn = 0, n even
1
c„ = -— , n odd V
JH7T
Thus
1 ±00 1
e(0 = z + V -ein7Tt
z n=±T±*,'.j'n7T
€i(ot €jUot-tan—la,)
1 +j(JD V1 + (JO2
where H(joo) is the a-c steady-state system function defined in Eq. 2.6-23.
Two limitations of the Fourier series method are that y(t) is expressed
as an infinite series rather than in closed form, and that only the steady-
state component is obtained. Because of this, the method is generally
t This closed-form solution is not obtained directly from the infinite series, but from
the classical solution of Example 2.12-3. e0 = di/dt, where i is given in Eq. 2.12-19.
Sec. 3.2 The Fourier Series and Integral 105
used only when the frequency spectra are desired, as in wave filters, and
not to calculate the output waveform as a function of time.
The Fourier series of Eqs. 3.2-6 and 3.2-7 can be rewritten with co = nco0.
'T! 2 i 'T/2
-n 1
, dt = — \ f(t)e-j0it dt
Aco T A00 J-T/2 2tt J-T/2
00
m= i ej(ot A to
C0=— 00 Aco
The symbol Aco denotes the spacing between lines in the frequency
spectrum, and it is equal to co0 = 2it/T.
A nonperiodic function of time can be obtained from a periodic one by
letting T approach infinity. Since Aco then approaches zero, the formerly
discrete frequency spectrum becomes a continuous one, containing all
possible frequencies. The quantity g(co) = cn/Aco is introduced because
cn itself vanishes as T approaches infinity. Taking the limit as T approaches
infinity and as Aco becomes dco,
I f00
g(*>) = r- ft*-’ dt
2rr J—oo
/* 00
m = g(o>)*iwt dco
J — CO
fit)
)\
1
L
L L
2 2
(a)
|G(to)|
|G(co)|
A
>■ co
0
(c)
Fig. 3.2-3
Example 3.2-2. Find the frequency spectrum of the rectangular pulse shown in
Fig. 3.2-3a.
'Ll 2
1 sin (coL/2)
G(co) - | - e-jcot dt =
-LI2 ^ coL/2
G(co), which is real because /(V) is an even function, is illustrated in Fig. 3.2-3b. In the
limit as L approaches 0, fit) becomes the unit impulse, and G(co) is shown in part (c)
of the figure. Thus the unit impulse contains all frequencies in equal strength.
This example demonstrates the principle of reciprocal spreading. The narrower the
function of time is made, the more its frequency spectrum spreads out, and the greater
is the bandwidth required to reproduce it faithfully. If the sharp corners in the function
Sec. 3.2 The Fourier Series and Integral 107
of time are smoothed, the high-frequency components in its frequency spectrum are
reduced. For example if, f(t) = (1V27r)e~t2/2, which is the Gaussian-type pulse of
unit area shown in Fig. \A-la, G(co) = e_c°2/2.
Example 3.2-3. Find the Fourier transform of the unit step function.
00
f* 00
t 00
G(co) €-ja)t dt =
Jo L-jnJ
cannot be evaluated, since the upper limit is not uniquely defined, i.e., the defining
integral does not converge. If the reader is persistent, he might note that
100
€-at€-jmt gt _
1
J5'[e_a< for t > 0] (a ^ 0)
'0 a +JW
which, as a approaches 0, reduces to 1 /jco. But this limit is not, mathematically, the
Fourier transform of the unit step function. To find the function of time corresponding
to G(co) = 1/yco, Eq. 3.2-9 gives
The purpose of this example is to point out that some of the very common functions of
time do not have Fourier transforms. The attempted use of the convergence factor
€~at suggests the heuristic derivation of the Laplace transform in the next section.
In contrast to the Fourier series, the Fourier integral of Eq. 3.2-11 gives
the complete response in closed form, provided that the system contains
no initial stored energy.
Example 3.2-4. For the circuit of Fig. 3.2-4o, plot the frequency spectrum of the
input and output, and find e2(t). Assume that there is no stored energy in the capacitance
at t = 0.
1
Ex(co) jajt dt
1 + jco
E2(co) = Ejco) 1 - 1
1 + jco (1 + jco)2
108 Transform Techniques
$E2(u)
(b)
Fig. 3.2-4
The frequency spectra are plotted in part (b) of the figure. The evaluation of
00
1
eft) = — (ho
2tt oo (1 +jcoy
involves a difficult integration. Tables of Fourier transforms do exist,1 and the use of
complex variable theory is often helpful. It can be shown by the methods of Section
3.6 that
' j CO
-ja>t
1
eft) = d(joj)
Irrj oo (1 + /to) ■
Sec. 3.3 The Laplace Transform 109
where the integration has been carried out with respect to the complex variable 5. The
details of the integration are not pursued further here, but the reader should realize
that the direct evaluation of the inverse Fourier transform is seldom easy.
where and cSfn-1 stand for the direct and inverse two-sided Laplace
transform.
This presentation is a heuristic one, illustrating the relationship between
the Laplace and Fourier transforms, rather than emphasizing mathe¬
matical rigor. In Eq. 3.3-1, the value of a = Re s is chosen to make the
integral converge, if possible. In Eq. 3.3-2, it is understood that the
integration is carried out with a within the same range that ensures the
convergence of Eq. 3.3-1. Since s is a complex variable, Eq. 3.3-2 involves
integration in the complex plane. A detailed discussion of this is postponed
until after a review of complex variable theory. Two examples of the use
of Eq. 3.3-1 are presented below.
110 Transform Techniques
for t < 0
/(') =
for t > 0, for real a
' 00 00
-is—a)t
zate-st rff
F(s) = _
_-(s - a)_|0 s — a
with a region of convergence given by a > a. The function in this example does not have
a Fourier transform if a is positive but does have a Laplace transform.
F(s) — — eate-st fa —
■ oo s — a — 00 s — a
that f(t) for t < 0 is of no concern and can hence be assumed to be zero.
I* 00
F(S)= /(t)*-1 dt = X[f(t)] (3.3-3)
Jo
A subscript I is not used, since the Laplace transform is assumed to be the
one-sided transform, unless otherwise indicated. The application of
Eq. 3.3 -3 to several common functions of time leads to the results in
Table 3.3-1:f
Table 3.3-1
Region of
fit) for t > 0 F(s) Convergence
u0(t) 1 G > — 00
1
u_ i(0 — a > 0
s
1
t G > 0
72
e~at
i
g > —a
s + a
P
sin fit G > 0
S2 + p2
s
cos fit G > 0
s2 + fS2
e-at sjn P
g > —a
(s + a)2 + p2
s + a
e~at COS pt g > —a
(s + a)2 + p2
e~a t f n—1
1
{n - 1)! g > —a
(s + a)n
n = 1,2,...
The expression for the inverse transform is the same whether the one¬
sided or two-sided transform is used.
■j (V-fj'oo
m = A F(sy‘ ds = y-’[F(S)] (3.3-4)
2ttJ Ja—jco
The integration must still be carried out with a within the range that
ensures the convergence of Eq. 3.3-3.
j A very extensive table appears in Reference 2. Tables of moderate size can be found
in several of the other references.
112 Transform Techniques
fi(t)
hit) hit)
Fig. 3.4-1
lim f(t)ekt — 0
t-+ oo
Examples of functions not of exponential order are e'2 and tl. A function of
nonexponential order may or may not have a Laplace transform.
Sec. 3.4 Properties of the Laplace Transform 113
correspondence between/(/) and F(s). This means that Table 3.3-1 may
be used to find the inverse as well as the direct transform, which in many,
but not all, cases obviates the need for Eq. 3.3-4. Note that the uniqueness
problem of Examples 3.3-1 and 3.3-2 no longer exists, because the two
functions given there do not have the same one-sided transform.
The use of residue theory in connection with Eq. 3.3-4 provides a
powerful and relatively simple method of finding /(/). As shown in
Section 3.7, the use of Eq. 3.3-4 in the one-sided transform yields a
function of time that is zero for t < 0, and that is continuous except for
possible singularity functions for t > 0. Thus FF~Y [l/Os + D] is identified
as the function f\{t) in Fig. 3.4-1, rather than f2(t) or /3(f).
The functions encountered in system theory often have discontinuities
at t — 0. It is usually convenient to define the value off(t) at t = 0 to be
the limit of f(t) as t approaches zero through positive values. This limit
is symbolized as/(0+), and for the function fx(t) in Fig. 3.4-1 has a value of
unity. Although it is not necessary to adopt this convention, most
engineers do so, for it leads to the fewest difficulties. Consistent with this
convention, and with the fact that only the response for t > 0 is usually
desired, the definition of the direct Laplace transform can be written as
Table 3.4-1 lists the most useful properties of the Laplace transform.
Proofs in most cases follow directly from Eq. 3.4-1 and can be found in
most references.4 Equations 3.4-2 through 3.4-7 are the basis for the
solution of fixed integral-differential equations by means of the Laplace
transform, as discussed in the next section. Equations 3.4-8 through
3.4-12 are useful in finding the direct and inverse transforms of given
functions.
Example 3.4-1. Find if-qi/U + 2)2].
Table 3.4-1
Equation
Property Number
r</ i
if = JF(J) -/(0+) (3.4-4)
if = 5wF(1y) — y(0+)
df
(3.4-5)
's i(0+) ^(0+)
H?)
if /(t) flfr (3.4-6)
5
F(s) /~1(0+)
if /(r) dr
I —CO 5 5
• 00
y«r
se F( r ) dr (3.4-10)
if = a F(as) (3.4-12)
/(i
a,
if Ad - W) dx = Fx{s)Fls) (3.4-13)
L Jo
1 Cx+j oo
nfiiOMO] = r-. F1(w)F2(s — w) dw
^‘Trj J x— j co
where w = x + jy, and where x must be greater
than the abscissa of absolute convergence for fx(t)
over the path of integration (3.4-14)
lim/(7) = lim sF(s) provided that the limit exists (3.4-15)
►0 s—*- co
Example 3.4-3. Figure 3.4-2 shows a periodic function of time fit), with period
T. Let fit) denote the first cycle only,/2(0 the second cycle only, etc. If
PT
fit)
Since
fit) = fiit) + fit) +/3(0 + '
ana since
fit) = fit - T)U_iit - T)
UO = fit - 2T)U_iit - IT)
Eiis)
Fis) = .-Ts
for Re s >0
1
116 Transform Techniques
This result is very useful when finding the response of a system to a periodic function.
Unlike the Fourier series method, both the steady-state and the transient response
may be found in closed form.5
The last two properties in Table 3.4-1 are known as the initial and final
value theorems, respectively. The meaning of the word analytic is dis¬
cussed in Section 3.6.
Impulses in f(t)
In many cases, F(s) is the quotient of two polynomials with real co¬
efficients. In such cases, the inverse transform can always be found from
Sec. 3.4 Properties of the Laplace Transform 117
P(s) P(s)
m = (3.4-17)
Q(s) (s - - sr+1) — sn)
where the denominator polynomial Q{s) has distinct roots ^r+1, . . . , sni
and a root sx which is repeated r times. If (2(X) is of higher order than
-P(V), then
Kr_ i K„
Hs) = + H-- +
(S ~ S1) (s - sf)' (s - sj
K r+l Kn
+ + +
S — S r+l S -
where
Kt = [(j - O' = r + 1, . . . , n)
Kr = [(5 - s.fFis)]^
d
Krl = (s - sf)rF(s)
_ds -1 S=Sl
and, in general,
1
Kr-k = (s - s1)rF(s) (k = 0, 1, 2, . . . , r — 1)
kl [_dsl S=S1
Once the partial fraction expansion has been determined, the inverse
transform of each term follows from Table 3.3-1.
Example 3.4-4. Find the inverse transform of
s3 + 1
F(s) =
S3 + 5
Since the denominator polynomial is not of greater order than the numerator poly¬
nomial, a preliminary step of long division is necessary.
1 (■\/2i2)e~j7T/4 (V2/2)ejV/4
F(s) = 1 + -s
s-J S+j
Using Table 3.3-1 and Eq. 3.2-5,
The expansion in the last example would certainly not be correct without
the 1 obtained by long division, since F(s) approaches unity for large
values of s. Whenever the order of the numerator equals or exceeds that of
the denominator, long division terms are needed to describe properly the
function for large values of s.
Letting y = — t,
r oo
G(a>) = f{—y)ea>y dy
00
dt = FR(-jco)
where fR{t) = /(—t) and is the original function reflected about the
vertical axis. If f(t) is nonzero for both positive and negative values of
time, it can be decomposed into two parts, each of which falls into one of
the two previous cases.
Example 3.4-5. Find the Fourier transform of the function shown in Fig. 3.4-3#,
given by
Let
m =m + m
where the last two functions are shown in parts (b) and (c) of the figure.
Finally,
2
G(co) = Giiw) + G2(co) =
w2+ 1
Sec. 3.5 Application of the Laplace Transform to Time-Invariant Systems 119
fit)
hit) f2R(t)
Fig. 3.4-3
Example 3.5-1. Find an expression for the response eft) for t > 0 in the circuit of
Fig. 3.5-1. The source eft) is a known function of time, and the current iL and voltage
ec are assumed to be known at the instant t = 0.
The circuit is described by the two loop equations
+ 3/2 + 2
/ h^ = 0
120 Transform Techniques
112 ec(t)
o
A
if
2h
112 S2 (t)
212
o
Fig. 3.5-1
Let Ifs), I2(s), and Efs) stand for the transforms of 4(0, 4(0, and eft), respectively.
Transforming the differential equations term by term yields the following algebraic
equations.
(25 + 3)4(0 ~ (25 + 2)4(5) = Efs) + 214(0+) - 4(0+)]
Note that
4(0+) — 4(0+) — 4(0+)
and
‘t
24-40+) = 2 lim 4(0 dr = ec(0+)
<--o+ ■ 00
Since 4(0 and 4(0 are the only unknowns, the two algebraic equations may be solved
simultaneously for 4(0- There results
2s2 + 25
eft) JS?-1 Efs)
452 + 95 + 6
The terms involving initial stored energy summarize the history of the
circuit for t < 0. Such terms frequently occur when a switching operation
is performed at the reference time t = 0, adding or disconnecting elements.
It is best to transform the differential equations immediately, rather than
first differentiating one of them, or solving them simultaneously. If the
equations are transformed immediately, the 0+ terms are always directly
related to the initial stored energy, and they vanish if the system contains
none.
d%V . i dy , o 1 dv ,
dt2 dt 2 dt
so
Ks + 2)
Y(s) =
s2 + 3s + 2 s + 1
and
2/(0 = *e-*
This answer agrees with the impulse response found in Example 2.8-1.
Alternatively, if v is a unit impulse occurring immediately before t = 0, the transform
of the right side of the differential equation is zero, but (dy/dt)(0+) = — £ and 2/(0+) =
£, as in Example 2.8-1. The complete transformed equation is
which gives the same result for y(t) as before. This alternative approach is not recom¬
mended because of the difficulty of finding the necessary initial conditions.
The assumption of no initial stored energy within the system is one that
is made throughout much of this book. It is not necessary in such cases
to write differential equations describing the system first and then trans¬
form them. Consider the typical electrical elements shown in Table 2.6-1.
When the element has no stored energy, the ratio E{s)jl{s) — Z(s) is the
simple algebraic quantity given in the last column. Each element can
therefore be characterized by an impedance Z(s), which is not a function
of voltage, current, or time. Thus, when elements are interconnected,
the rules for d-c and a-c steady-state analysis carry over to Laplace
transform analysis. The relationship between the transformed output
and input can be found by the solution of algebraic equations.
Example 3.5-3. Find an expression for e2(t) for / > 0 in the circuit of Fig. 3.5-1,
assuming that there is no initial stored energy.
The circuit is redrawn in Fig. 3.5-2, each element being labeled with its impedance,
and transformed voltages and currents being used. Any method from d-c circuit theory
122 Transform Techniques
2/s
Ei(s)
may be applied. Perhaps the easiest is to write a single node equation at node 3.
1 1
Efs) 1 + —-: + = E x(s)
2s T 2 1 + 2/s
or
(2s + 2)0 + 2)
EAs) =-EAs)
J 452 + 95 + 6 '
Then
1
E‘2 0) =
1+2/5
Efs)
or
2s2 + 2s
e2(t) = Efs)
4 s2 + 95 + 6
There is a direct relationship between the input v(t) and output y(t)
for a linear system with no initial stored energy. The system function
(or “transfer function”) of a fixed system is defined as the ratio of the
transformed output to the transformed input.
H(s) =
r(s) (3.5-1)
V(s)
or
2/(0 = Se-i[V(s)H(s)] (3.5-2)
and system functions defined in this section. Recall that H(s)est is the
forced or particular integral response to the input est. From Eqs. 2.6-21
and 2.6-22,
B(s) yp(t)
H(s) = (3.5-3)
-40) /
v(t)=e
st
where A(s) and B(s) are formed from the operators A(p) and B(p) in the
differential equation
Mp)y(t) = B(p)v{t) (3.5-4)
s2 + 3s + 2 S + 1
which the reader will recognize as the transform of the impulse response.
The use of Eq. 3.5-2 constitutes the same basic three-step procedure
used for the complex Fourier series and the Fourier integral. The input
function is transformed (this time into the complex frequency domain)
and is then multiplied by the system function. The resulting transformed
output is then converted back to the time domain. In fact, when s is
replaced by jo) in the system function, H(jco) is the same system function
that was used with the Fourier series and integral. From Eq. 2.6-23,
Thus the inverse transform of the system function is the impulse response.
This is one reason for using the symbol h(t) for the impulse response.
Similarly, using Eq. 3.5-2 with v(t) = U_±(t) and y(t) = r(t),
(3.5-7)
Example 3.5-4. Solve Example 1.6-3 by use of the Laplace transform. The example
considers n identical amplifiers in cascade, as in Fig. 3.5-3, each having the step response
r(t) = —Ae~t,RC for t > 0.
Since
—A
s + 1 IRC
124 Transform Techniques
En(s)
Fig. 3.5-3
Assuming that the individual amplifiers are not loaded down by the presence of suc¬
ceeding stages, the system function for the combination is
1 / —As
s \s T 1 /RC/
The inverse transform is6
j n^Hn~ 1 )\(-t/RC)k
(-A)ne~tlRC
£i(n - k -1)!(*!)2J
agreeing with the result given in Example 1.6-3.
This result is identical with the convolution integrals of Eqs. 1.6-2 and
1.6-4 for nonanticipatory systems with v(t) = 0 for / < 0. Also,
H(s)
y(t) = J?-1 sK(s)
s J
The inverse transform of the last factor in the brackets is r(t) by Eq. 3.5-7.
By Eq. 3.4-4,
y-'lsFfs)] = —
dt
ifr(0+) = 0. Using Eq. 3.4-13,
Since v(t) is regarded as zero for t < 0 in the one-sided transform, i>(0+)
is the size of a discontinuity at the origin. Then dv(A)/dA will have an
impulse of value y(0+) at A = 0. To be consistent with the definition of
the one-sided transform in Eq. 3.4-1, however, Eq. 3.5-9 should more
properly be written
where a and a> are the real and imaginary parts of 5, respectively. As for
any complex quantity, values of 5 can be represented graphically in a
complex plane, shown in Fig. 3.6-1. It is sometimes convenient to think of
an “extended s-plane” as the surface of a sphere of infinite radius, with the
126 Transform Techniques
« (imaginary axis)
A
s-plane
cr (real axis)
Fig. 3.6-1
R and X are the real and imaginary parts, respectively, of F(s), and they
are real functions of the two real variables a and co.
Analytic Functions
oo
(R da — X dco) + j (X da + R dco)
Jc c
The values of a and co are related by the contour along which the integra¬
tion is carried out.
A contour integral of F(s) between two fixed points in the 5-plane is not,
f The phrase “analytic at and in the neighborhood of a point in the 5-plane” is more
correct mathematically, but it is also more cumbersome.
128 Transform Techniques
In Fig. 3.6-2,
(j> F(s) ds = F(s) ds — F(s) ds
F(s) ds = 0 (3.6-6)
'c
A corollary to this theorem is that
in Fig. 3.6-2, if F(s) is analytic on and between the two paths. Another
corollary deals with multiple connected regions. In Fig. 3.6-3a, F(s) is
assumed to be analytic on the closed contours C, C1? and C2, and in the
shaded region R. It is not, however, necessarily analytic inside Cx and C2.
In Fig. 3.6-3b, cuts labeled C3 through C6 are constructed so as to give
a simply connected region. The distance between corresponding ends of
C3 and C4, and also C5 and C6, is assumed to be infinitesimal. By Cauchy’s
theorem.
f
1C3
ra
e
Cs) ds —
J Ci
r%
F(s) ds
F(s) ds + F(s) ds
/c. Cs
r
/%
F(s) ds F(s) ds = 0
C2 kJ Ce
CO co
A C
(a)
Fig. .6-3
Sec. 3.6 Review of Complex Variable Theory 129
since F(s) is analytic on and inside the closed contour formed by the sum
of these individual contours. By Eq. 3.6-7, the integration along C3 and
C4, and also along C5 and C6, cancel, so
Taylor’s Series
and the first k — 1 derivatives are zero at 5 = 50, the first k terms in Eq.
3.6-11 are missing, and F(s) is said to have a zero of order k at s0.
A uniqueness theorem states that, if two power series represent F{s)
in the neighborhood of 50, then they must be identical. The coefficients
may therefore be found in any convenient way, and not necessarily from
Eq. 3.6-12. In the examples which follow, long division is used.
The uniqueness theorem of the previous paragraph leads to other useful
results. If a series representation of F(s) is valid in any arbitrarily small
region, it is the unique representation wherever it converges. Also, if an
analytic function is specified throughout any arbitrarily small region, it is
then uniquely determined throughout the entire 5-plane. The last state¬
ment is known as the principle of analytic continuation.
Example 3.6-1. Expand F(s) = 1/(5 + 1) in a power series about 5 = 1, i.e., in powers
Of (5 — 1).
By long division,
1 1 1
- (5 - 1) + - (5 - l)2 - — (5 - l)3 + * • *
5 —f- 1 2 —(5 — 1) 2 4 8 16
CO
A
Circle of
Fig. 3.6-4
The same series results if Eqs. 3.6-11 and 3.6-12 are used with s0 = 1. The function
F(s) is analytic except at the point 5 = — 1, shown by the cross in Fig. 3.6-4. The circle
of convergence indicates that the series converges for \s — 1| <2.
5 1 + (5 — 1) 3 3
5-3 -2 + (5 - 1)
Sec. 3.6 Review of Complex Variable Theory 131
which converges for Is — 1| <2. Dividing each term of the series by (s — 1),
0 - l)(s - 3) s — 1 4 8
Laurent’s Series
The hrst part of the series is known as the ascending part; the second part,
the principal or descending part. The series converges between two circles
of convergence, both centered at s0. F(s) must be analytic between these
circles, which are labelled Cx and C2 in Fig. 3.6-5. If there is an “isolated
singularity” of F(s) at s0, the circle C2 may shrink to infinitesimal size, and
the Laurent series then represents F(s) in the vicinity of s0. As discussed
later in this section, the coefficient b_x is particularly important and is
called the residue of F(s) at s0.
In the use of Eq. 3.6-14, the contour C may be any closed contour
between C1 and C2, shown in Fig. 3.6-5. Since this equation is difficult
to evaluate, however, the coefficients in Laurent’s series are normally
Fig. 3.6-5
132 Transform Techniques
found by some other means. Any convenient method may be used, be¬
cause the representation of F(s) by a series in powers of (5 — s0) in any
given region is unique. Example 3.6-2 is an example of a Laurent series,
where C2 is arbitrarily small, and Cx is a circle of radius just under 2.
Classification of Singularities
Singularities of F(s), also called singular points, are points in the 5-plane
at which F(s) is not analytic. If nonoverlapping circles, no matter how
small, can be drawn around each singular point, the points are called
isolated singularities. The function F(s) = (5 + l)/[s3(s2 + 1)] has
isolated singularities at 5 = 0, -f-j, and —j. The function F(s) = 1/
[sin (77/5)] has isolated singularities at 5 = 1, J, . . . , but a nonisolated
singularity at the origin. Fortunately, the commonly encountered
functions have only isolated singularities. The reader should remember
that F(s) can be represented by a Laurent series in the vicinity of every
isolated singularity.
An isolated singularity is classified further by examining the Laurent
series written about it. If the principal or descending part of the series has
an infinite number of terms, the singularity is called an essential singularity.
Otherwise, the singularity is called a pole. The order of the pole is equal
to the number of terms in the principal part of the series. If F(s) has a
pole of order n at 50, then there are no terms in the principal part of the
power series for (5 — s0)nF(s), so that (5 — s0)nF(s) is analytic at 50.
around the individual circles, as in Eq. 3.6-8 and Fig. 3.6-3. Applying
Eq. 3.6-15 to each individual circle,
where the summation includes the residues at all the singularities inside C.
g(s) -n+l
F{s) = n
= g(s0)(s - s0) n + g'(s0)(s - s0y
(s - So)
~g(n ^(Sp)' £(n)(s0)
+ (s - s0) 1 + + •
.(n - 1)!J L n
The last expression is the Laurent series for F(s), so the residue at s0 is
1 r dn~x
b-i = — (s - so)"F(s) (3.6-17)
{n - 1)! ds -1 S=S 0
P(s)
F(s) = (3.6-19)
Q(s)
where both P(s) and Q(s) are analytic at s0, and where P(s0) ^ 0. If F(s)
has a first order pole at s0, then Q(s0) = 0, but Q'(s0) ^ 0. Writing a
134 Transform Techniques
Taylor series for both P(s) and Q(s) and carrying out the indicated long
division,
P(s0) + P'(s„)(s -$„) + ■■■
f(s) =
Q'(so)(s - So) + ^ <2"(s0)(s - S0)2 +
P(Sq)
0 - s0) 1 +
QXso)
The residue at 50 is
P(sp) '
b -1 =
(3.6-20)
QXs0)
This approach becomes cumbersome when applied to higher order poles.
When the Laplace transform F(s) is known, the function of time can be
found by Eqs. 3.3-2 and 3.3-4, rewritten below.
f{t) = — lim
277j R-* oo Jc—jR
P F(s)€s< ds (3.7-1)
As this equation applies equally well to both the two-sided and the one¬
sided transform, it is convenient to treat them together.
Recall from Section 3.3 that the defining equation for the direct trans¬
form F(s) converges only for certain values of <r, i.e., only within a certain
region of the 5-plane. Using the principle of analytic continuation, how¬
ever, this is sufficient to uniquely define F(s) throughout the entire 5-plane,
except at the singular points. Since the factor est is analytic throughout
the entire finite 5-plane, the function F(s)est can be integrated without
difficulty along any path that does not include any singularities of F(s).
If the concept of an extended 5-plane mapped on the surface of an infinite
sphere is used, the point at infinity must be avoided, since est has an
essential singularity there. In the application of Eq. 3.7-1, a semicircular
detour could be made around the point at infinity. An easier interpretation
of Eq. 3.7-1, however, is to consider only the finite 5-plane, and to apply
the limiting process of R —► oo after the integration has been carried out.
This is the procedure that is followed here.
In Section 3.3, it was indicated that the path of integration in Eq. 3.7-1
is restricted to values of a for which the direct transform formula con¬
verges. In the case of the two-sided Laplace transform, the region of
convergence must be specified in order to uniquely determine the inverse
Sec. 3.7 The Inversion Integral 135
co
A
E D
V 0
c
< >
B
H
G A
Fig. 3.7-1
00 JDEFGA
for t > 0, and
lim F(s)est ds = 0
R^oo Jdiia
for t < 0. In this event, path ABD may be changed to ABDEFGA for
136 Transform Techniques
t > 0, and to ABDHA for t < 0. Using the residue theorem of Eq.
3.6-16,
1 1
f(t) = — lim O' F(s)est ds
2irj R-+co Jabdefga
i
m lim O F(s)est ds
2rrj r -* 00 jABDHA
In using Eqs. 3.7-2 and 3.7-3, the residue formulas developed in Section
3.6 prove helpful. For a pole of order n at s0, the residue of F(s)est is
1 d n—1 st
n-
~(s~ so)nF(s)< (3.7-4)
(n- 1)! .ds s=so
f The theorem as stated is really unnecessarily restrictive. For example, Eqs. 3.7-2
and 3.7-3 are valid if F(s) is any meromorphic function (the ratio of two functions that
are analytic throughout the finite s-plane) that approaches zero uniformly as \s\ co.
If there is ever any doubt, Eqs. 3.7-2 and 3.7-3 may be used formally to find a function
of time. If the transform of this answer is found to be the original F(s), then the validity
of the answer is established.
Sec. 3.7 The Inversion Integral 137
co
Since Theorem 3.7-1 is satisfied, Eqs. 3.7-2 and 3.7-3 may be used. For t > 0,
fit) = [residue of F(s)est at 5 = 0]
= t + 1
s=0
For t < 0,
fit) = —[residue of Fis)est at 5 = 1]
The complete function is shown in Fig. 3.7-3. The region of convergence in this example
indicates that the two-sided Laplace transform must have been used in obtaining F(s).
Poles to the left and right of this region yield, respectively, the components of /(/)for
positive and negative time.
fit)
Fig. 3.7-3
138 Transform Techniques
Example 3.7-2. F(s) = (s + tf)/[0 + a)2 + /?2] = 0 4- a)/[(s + a — jP)(s -+■ a + j(3)].
If the region of convergence is given by o > —a, find f(t).
Equations 3.7-2 and 3.7-3 may again be used. For t > 0,
(s + a)est (5 + a)est
+
_s + a + jfj_ s=-a+jfi _s + a — jfj_
For t < 0, f{t) = 0, since there are no singularities inside path ABDHA in Fig. 3.7-1.
—WVl/wv --WVW-WV—AA^-t-vW-vVA-f-vW-^1
R/2 w* R/2 R/2 R/2 R/2
Characterizing each element by its impedance, the transformed equation for the
rth mesh is
1 1
— Ir-i(s) +(/?-!—— ITC5)-t;T+i(A — 0
sC sC sC
By the method of Example 2.12-1, the solution of the difference equation, subject to the
boundary conditions above, is
C sinh [(m — r)0]
Ir(s) =
(sinh 0)(cosh md)
where cosh 0 = 1+ (RC/2)s. The hyperbolic functions are analytic for all finite values
of the complex variable 6. Whenever sinh 0 = 0, [sinh (m — r)0]/sinh 0 is finite, so the
only poles of Ir(s) occur when cosh md = 0. Letting 0 = a + y/3,
This expression is zero only when m(3 = ±(2k — l)(rr/2) and men — 0, where k =
1, 2, 3, ... . The poles are at
.(2k - 1)tt-
»- ±) 2m
(2k — 1)77
cosh 0 = cos
2m
2 (2k - 1)77
j = — 1 + cos
RC 2m
The ± sign is not needed because cos (—<f>) = cos </>. Also, after k = m, the values of
5 in the last equation are repeated. Thus there are m poles, as might be expected from
the fact that Fig. 3.7-4 has m energy-storing elements. The poles lie on the negative
real axis of the j-plane between the origin and —4/RC.
The integral of F(s)est along path DEFGA in Fig. 3.7-1 can be shown to vanish for
t > 0 as R —00, so f(t) is the sum of the residues of F(s)est at the m poles. Define
dQ(s) dQ dd RC
-= —--= [(sinh d)(m sinh md) + (cosh 0)(cosh md)]
ds dd ds 2 sinh 0
At the poles, where cosh md = 0, and sinh md = j sin [(2k — \)(tt/2)] = y( —l)fc+1,
dQ , RCm\
ds s=s.
= ./(-Dfc+1(— J
The fact that the last expression is not zero indicates that the poles are of the first order;
hence Eq. 3.7-6 may be used.
SF-
J2(52 + J + 1)
140 Transform Techniques
The procedure most commonly used is to find, in the normal manner, that
1 V3 77
se- — t — 1 H-- e-</2 cos — t H—
_s2(s2 + s + 1) V3 2 6,
V3 V3
JV- t — 2 H-- e <f-1)/2 COS — t--1- U^{t - 1)
S2(S2 +5+1) V3 \ 2 2 6,
which is the previous function of time shifted along tjie time axis one unit in the positive
t direction.
If, however, the inversion integral were to be used directly,
i -s(t-l)
m= 2-nj
ds
ABD s\s + 1 ~j)(s + 1 + j)
where path ABD in Fig. 3.7-1 passes to the right of all the poles.
:sU-l)
ds
s2{s + 1 — j){s + 1 +/)
would be expected to vanish along path DEFGA only for t — 1 > 0, and along path
DHA only for t — 1 <0, since the real part of s{t — 1) should be negative. The reader
should verify that this is indeed the case.
2 /V 3 V3 77 v
= t — 2 4-- e <( 1)/2 cos I — t-1-
v 3 \ 2 2 6/
co
A
Branch cut
e <j
Fig. 3.7-5
still applies, but Fig. 3.7-1 must be modified, as shown in Fig. 3.7-6. By the residue
theorem.
since no singularities are enclosed. For t > 0, it can be shown that the second and
fourth integrals vanish as R approaches infinity. Then
which vanishes as r approaches zero. On the straight line FG, let 5 = — u, sA = jiM,
and ds = —du, where u and iM are real, positive quantities. Note carefully that the
branch cut requires V — 1 = j and not —j for this path. Then
' oo ut
0 1 1
F(s)est ds — — du
FG oo juX/2 J Jo u
CO
142 Transform Techniques
1
V nt
The substitution y = ut was made, and the evaluation of the last integral follows either
from the definition of the gamma function or from a standard table of definite integrals.
Examples of finding the inverse transform of double-valued functions when the path
shown in Fig. 3.7-6 encloses poles of F(s) are discussed in Reference 10.
If the integral along the infinite semicircles DFA and DHA vanishes,
the contour ABD may be replaced by either of the two closed contours in
the figure. Then, by the residue theorem,
&mm\ = 2 [residues of F1(w)F2(s — w) at the poles of /^(w)]
= — 2 [residues of F1(\\ )F2(s — w) at the poles of F2(s — u’)]
(3.7-8)
Example 3.7-6. Find 3F[teat] by the lise of the complex convolution theorem.
If fiil) = t and f2(t) = eat, Eq. 3.7-8 gives
1
ST[teat] = residue of at w = 0
w2(5 — w — a)
1 1
dw\s — w — a, w =0 is ~ a)‘
Alternatively,
1
JT[teat] = residue of at w = s — a
w2(s — w — a)
1
_w j Hj=s_a is - a)‘
Poles and zeros are defined in Section 3.6. In brief, poles and zeros of
F(s) are values of s for which 1 jF{s) and F(s), respectively, become zero.
Suppose that F(s) can be written as the quotient of two factored parts.
co
A
1 o
Fig. 3.8-1
The output of a fixed linear system initially at rest is given by Eq. 3.5-2,
repeated below.
y(t) = (3.8-2)
Sec. 3.8 The Significance of Poles and Zeros 145
The poles of V(s) depend only upon the input and give rise to the forced
response, which is the particular solution in the classical language of
Chapter 2. The poles of H(s) depend only upon the system and yield the
free or natural response. This is identical with the complementary
solution of Chapter 2, with all the arbitrary constants evaluated. Since
the size of the terms in the free response depends upon the poles and zeros
of both V(s) and H(s), the arbitrary constants depend upon both the input
and the system, as stated in Section 2.6.
Example 3.8-1 Find the output voltage e{t) if the circuit of Fig. 3.8-2 has
no initial stored energy, and if i(t) = (1 + sin t)U_ft).
Fig. 3.8-2
1 s
Z(s) =
2/s + 5 + 3 (s + DO + 2)
_ 1 1 _ 52 + 5 + 1
5 52 + 1 S(S2 + 1)
Hence
s2 -F 5 + 1
E(s) =
(s + 1)0 + 2)02 + 1)
Then
1
e(t) = &~l[E(s)] = - e-* ~ - e-2t + —= cos
2 VlO
The first two terms in the output constitute the free response and are produced by the
poles of Z(s). The last term is the forced response, produced by the poles of I(s) at
s = ±j. There is no term in the forced response due to the pole at the origin, because
it is cancelled by a zero of Z(s). This can be physically explained from the circuit by
noting that the inductance acts like a short circuit to the d-c component of the source.
In this example, the free and forced responses are the transient and steady-state com¬
ponents, respectively. This is always the case if H(s) has only left half-plane poles, and
if the input does not vanish as t approaches infinity.
Stability
Since SP-^His)] = h(t), the poles of H(s) determine the form of the
impulse response. Because an impulsive input simply inserts some energy
146 Transform Techniques
into the system instantaneously, the nature of h{t) indicates the stability
of the system. The system is unstable if h{t) increases without limit as t
approaches infinity. Thus a system is unstable if H(s) has poles in the right
half-plane or higher order poles on the imaginary axis. If there are first
order poles on the imaginary axis, the free response contains oscillations
of constant amplitude. Such a system is usually, but not always, con¬
sidered to be stable. If H(s) has only left half-plane poles, the system is
unquestionably stable. These conclusions agree with those reached in
Section 2.6. By Eq. 3.5-3, the poles of H(s) are the roots of the character¬
istic equation A(s) = 0.
Previous sections of this chapter have discussed in some detail the use
of transform techniques and the system function H(s) in the analysis of
fixed systems. This section investigates the application of these techniques
to time-varying systems.
The first of the two basic transform methods for fixed systems was to
transform the system’s differential equations into algebraic equations.
Although any differential equation may be transformed term by term,
an algebraic equation results only if the coefficients of the differential
equation are constants. Otherwise, another differential or integral
equation results. Equations 3.4-9 and 3.4-10 are useful in solving homo¬
geneous differential equations whose coefficients are polynomials in t.
Example 3.9-1. Solve the equation
d2y dy
t —- -1—1— (t — \)y = 0
dt2 dt
when 2/(0+) — 1.
Transforming this equation term by term, using Eq. 3.4-9, yields
d dY(s)
- - [s*Y(s) - s - 2/(0+)] + [sY(s) - 1] + —+ Y(s) = 0
ds ds
Simplifying,
dY{s)
(s + 1) —— + Y(s) = 0
ds
which is a first order differential equation in Y(s). This can be easily solved by the
separation of variables, or by the method of Section 2.5, giving
K
Y(s) =
s + 1
Then
2/(0 = K*~*
Sec. 3.9 Application to Time-Varying Systems 147
2/(0 = €"*
Example 3.9-2. The differential equation of the previous example is changed slightly
to read
d2y dy
+ -t* + ty = 0
1 dt2 at
with 2/(0+) = 1.
The reader will recognize this equation as Bessel’s equation of zero order, whose
solution is y = J0(t). The transformed equation is
d dY(s)
- - [s2Y(s) - s - 2/(0+)] + |sY(s) - 1] - -fd = 0
ds ds
Simplifying,
dY(s)
(s2 + 1) + jT(0 = 0
ds
whose solution is
K
Y(s) =
Vs2 + 1
The constant K, needed to satisfy the condition 2/(0+) = 1, can be found by the initial
value theorem of Eq. 3.4-15.
sK
lim —— = K
S->00 Vs2 + 1
so K — 1, and
1
2/(0 = &-1
vV + l
Although the inverse transform appears in standard tables, note instead that, since the
solution is known to be J0(t),
-2Vo«] = — ■■ ■
Vs! + 1
The last example gives the easiest method of finding the Laplace
transform of the Bessel function. In general, the use of Eq. 3.4-9 yields
a differential equation in Y(s) of order equal to the degree of the highest
polynomial in t. It should be emphasized that the resulting equation
cannot always be easily solved, and that the preceding examples are not
typical of the difficulties encountered.
Each term in a differential equation with variable coefficients is the
product of one known and one unknown function of time. In trans¬
forming such a term, Eq. 3.4-14 can be used. The result is a complicated
integral equation in y(T). Since the resulting equation is very rarely easier
to solve than the original one, this method is not pursued further.
At this point, one might seek other integral transforms helpful in
solving those differential equations that are not amenable to the Laplace
148 Transform Techniques
(3.9-17)
3 u.
o 0 m NO
ON 1-H *—< ’—1 I
£ I 1 1 I
On On ON ON On On ON On
m rb co rb rb rb m rb
C/2
txO
3
‘>7 03
u
3
>■
03
P
3 t-
o
(3.9-8)
i
c7 rb 3" 77 Co t-~
- £ On
■
ON ON On ON
Table 3.9-1
On ON
si m rb rb rb rb C*~i
w z
= j^-i[r(5)//(5)]
3
o
Laplace transform
o
03 3
Response by the
C/3 3
3 C/3 -0 CL 3
o 3 C/3
3 C/3 3
03
Oh _o <3 03 O
C/3 s- 4—*
03 +-> c/3
O 3
I— 0
C/3 3 O C/3 <U
03 O s_ 3 C/3
C
4—>
C/3 Cu 5J0 .0 O 3 3 O
3
s— 03
03 -*—»
L—» c CL O Cl
X<D
im
3 3 Cl
CL CL 3 C/3
—s 03 L— 03
E GO U,
Sec. 3.9 Application to Time-Varying Systems 149
transform. This has been done, and a number of other transforms are
available for certain special cases. The best-known of these are the Mellin
and Hankel transforms.11 General integral transforms are discussed
further in Section 3.10.
The second basic transform method for fixed systems was based upon
the system function H(s). For systems with no initial stored energy, the
response to an arbitrary input is given by
Zadeh has shown that the system function concept can be extended to
time-varying systems.12 When the characteristics of a system are changing
with time, the system function must be an explicit function of t as well as s.
One would expect that the response of a varying system with no initial
stored energy would be given by
y(t) = 5)]
In order to understand how the time-varying system function results from
a generalization of H(s), first examine Table 3.9-1. Equations 3.9-1
through 3.9-8 summarize the three principal methods of characterizing
fixed systems. They follow directly from Eq. 2.2-6, the definition of the
impulse response, Eq. 1.6-2, Eq. 1.6-4, Eq. 3.5-6, Eq. 3.5-6, the discussion
following Eq. 2.6-20, and Eq. 3.5-2, respectively. For varying systems,
Eqs. 3.9-9 through 3.9-13 follow directly from Eqs. 2.2-7, 1.7-1, 1.7-1,
1.7-2, and 1.7-4, respectively. Equations 3.9-14 through 3.9-17 are derived
in this section.
In the differential equations 3.9-1 and 3.9-9, p = djdt. While the oper¬
ators A and B are functions only of p for fixed systems, they are explicit
functions of both p and / for varying systems. For example,
Since the exponential input was assumed to exist for all t, H(t, s)est really
represents only the forced or particular integral response. A similar
comment applies to Eq. 3.9-7, as is inferred by Eq. 3.5-3.
From Eq. 3.9-12, the response to an arbitrary input is
r 00
y{t) = v(X)h(t, X) dX
J — 00
K(s)€sA ds h(t, X) dX
c
where C is the contour of integration used for the evaluation of the
inversion integral of Section 3.7. Assuming that the order of the two
Sec. 3.9 Application to Time-Varying Systems 151
d2y dy
t2 —• +41-b 2y = v
dt2 dt
Write
1 1
2/(0 = ^-l[LVV(s)yH{Vt,, 0171 = t%
- ^_1 + ])2j
d
7s (s + 1)2J =0
and at ^ = — 1 is
= (/ + 2)€-‘
s=—1
Thus
e -t
dy
+ 4t -f + 2y = v
dt
Then
1 i
m, s) = - &ir\ =
t2S2
bk(t)pkest = bk(t)skest
so
B(p, t)est = estB(s, t) (3.9-18)
pVs) = 2 (kJ(pk-rf)(p'g)
'k'
where I ) represents the binomial coefficients, i.e.,
(c + df = r=o
^ (/v I c k-r dr
\ r>
Letting/ = H{t, 5) and g = est
k /k
Pk[H(t, s)es<] = 2 (l) [pk~rH(t, 5)][s^csi]
r=0 \'/
Thus
A(p, s)est] = estA(p + s) (3.9-19)
The methods of Sections 2.7 and 2.8 are often helpful in solving this
equation.
d2y dy
t2 — + 4/ - + 2y = v
dt2 dt
In operator notation,
(t2p2 + 4tp 3- 2)y = v
Using Eq. 3.9-20,
[t2(p + s)2 + 4 t(p + s) + 2\H(t,s) = 1
or
[t2p2 + {1st2 -f- 4t)p + (t2s2 + 4ts + 2)\H(t, s) = 1
does satisfy the last equation and does agree with the result of the previous example.
For this particular system, solving the differential equation in H(t, s) is more difficult
than solving the original differential equation for the impulse response. For other
systems, however, it may be simpler.
The reader should realize that the H(t, s) found in the last example is
only the forced or particular integral solution of Eq. 3.9-20. From the
derivation of Eq. 3.9-16, however, this is exactly what is desired. Al¬
though a different H(t, 5) results if a complementary solution is added, it
can be shown that the response calculated by Eq. 3.9-17 is not affected.12
In many problems, the exact determination of H(t, s) is so difficult that
recourse to approximation methods is necessary. Standard approximating
techniques may be used when the parameters either vary slowly with time
or have a small variation compared to their mean value.13
material of this section, the reader may wish to review Section 1.3.
Equations 1.3-3 and 1.3-4 express the input and output of a system
initially at rest as
v(t) = I a(X)k(t, X) dX
c
(3.10-1)
y(t) = I a(X)K(t, X) dX
Jc
If X is a real variable, the integration is in general from X = — oo to + go.
If X is a complex variable, the integration is over a contour C in the com¬
plex plane. k(t, X) represents the family of elementary functions into
which v(t) is decomposed. The spectral function a(X) is a measure of the
relative strength of the elementary functions comprising v(t). K(t, X)
is the system’s response to k(t, X).
In order for Eq. 3.10-1 to be useful, there must be a simple method of
finding a(X) for an arbitrary input. Because of the linearity of Eq. 3.10-1,
one would expect a(X) to be expressible in the following form.
OO
00 (3.10-4)
v(t) = a(X)U0(t — X) dX
J — oo
’ oo
= v(X)U„(l - A) cO
J — 00
1
y{ 0 = a(X)H(t, X)eudX
2Trj Jc
= ^Ir1[a(2)/7(t, 2)] (3.10-7)
where a(X) = u[v(t)\, agreeing with Eq. 3.9-17.
For any choice of elementary functions k(t, 2), the characteristic function
K(t, 2) can be found once the impulse response h(t, 2) is known. Using
Eq. 3.10-6 with v(t) = k(t, z) and y(t) = K(t, z),
oo
Conversely,
For the special case of k{t, X) = ex72vj and k~\X, t) = these two
equations should relate the system function H(t, s) of Section 3.9 to the
impulse response. For this special case, K(t, z) = {\j2zTj)H(t, z)ezt by
Eq. 3.9-16. Substituting these values into Eq. 3.10-8 gives
Letting £ = t — k,
h(t,z) = — X^'dX
IttJ Jc
= — f H(t, X)ew-Z) dX
2ttj Jc
Then
= &ll-1[H(t,X)] (3.10-11)
The spectral function a(K) for the input v(t) is often denoted by V(X).
Using this notation, Eqs. 3.10-2 and 3.10-1 become
oo
V(X) = v(t)k~\x, t) dt
-oo
v(t) = V(X)k(t, X) dX
c
= Y(X)k(t, X) dX
where Y{X) is the spectral function or transform of the output y(t) with
respect to the set of elementary functions k(t, X). The relationship
between Y(X) and V(X) becomes particularly simple if the characteristic
function K(t, X) happens to have the form K(t, X) = H(X)k{t, X). When
this occurs, the elementary functions k(t, X) are said to be “eigenfunctions”
of the system under consideration.
response to k(t, X)
H(X) =
k(t, X)
and is called the system function with respect to k(t, X). In such a case,
V(X)H(X)k(t, X) dX
V(X) = I r(t)/c_1(2, t) dt
J — co
y(t) = Y(X)k(t, X) dX
Jc
This specifically describes the three steps associated with the Fourier and
Laplace transforms: the direct transform, multiplication by the system
function, and the inverse transform.
The forced response of a fixed linear system to eu always has the form
H(X)eu, as shown in Eq. 2.6-22.| For such systems, the previous equations
t Zadeh points out that the statement is also true for a certain class of nonlinear
systems.15
158 Transform Techniques
become
I* oo
V(X) = \ v(i)<Tlidt = S?u[v(t)]
J — o0
where the lower limit becomes zero if ift) = 0 for t < 0. The two-sided
and one-sided Laplace transforms of Eq. 3.11-1 are
oo
F„*(s) = 2 v(nT)e~nTs
n=— oo
oo (3.11-2)
F*(s) = 2 v(nT)e~nTs
n=0
Sec. 3.11 The Z Transform 159
2 = esT (3.11-3)
Vu(z) = J v(nT)z~"
n—— 00
„ (3.11-5)
V(z) = ^v(nT)z~n
n=0
where T is a known constant. Since T(2) depends only upon the value of
f(t) at the instants t = nT, it can be equally well regarded as the Z trans¬
form of either f(t) or
An alternative interpretation is based upon the fact that the exponent
of 2 in Eq. 3.11-6 denotes a particular element in the infinite sequence
. . . ,/( — . . . ,f{nT), .... The variable 2 is often regarded
as an “ordering variable” for the infinite sequence.
Equation 3.11-6 defines F(z) only for those values of 2 for which the
infinite series converges. For other values of the complex variable 2,
F(z) is defined by the principle of analytic continuation. Whenever ^(2)
is written in the infinite series form of Eq. 3.11-6, the values off{nT) can
be seen by inspection. If F{z) is given in closed form,f{nT) can be found
by the techniques of the next section.
160 Transform Techniques
|0 for / < 0
^ ieat for t > 0
By Eq. 3.11-6,
oo oo aT
f(z) =y tmTz-n =y —
n=0 n=0
This infinite series converges for
aT
aT
< 1, i.e., |z| >
00 00
= - 2 (e~aTz)n+1 = -e-aTz^ (e-aTz)n
n=0 n=0
The infinite series converges for \z\ < eaT and can be written in closed form as
F(z) = -e~aTz
1 — e z z — e a!1
or, equivalently,
F(z) = {J?[f*(t)]}€,r=z (3.11-9)
Sec. 3.11 The Z Transform 161
Table 3.11-1
Region of
m F(z) Convergence
z
£/_i(0 z - 1
1*1 > 1
Tz
t 1*1 > 1
(z - l)2
T2z(z + 1)
t2 \z\ > 1
(2 - 1)3
Z
t,at 1*1 > €
z„ — e aT
z sin (IT
sin [It 1*1 > 1
z2 — 2z cos pT + 1
z(z — cos 0D
cos [It 1*1 > 1
z2 — 2z cos [IT + 1
The presence of the last term compensates for the fact that n — 0 appears
in both of the summations. Letting m = —n,
co oo
= + &\f( o] -m (3.H-10)
nntiT)] = 2 /(«>-»
71=0
which does not depend upon T. This theorem is useful in finding the in¬
verse transform by partial fractions, particularly if the available tables
assume T = 1.
Table 3.12-1 lists the most useful properties of the Z transform. Equa¬
tions 3.12-1 through 3.12-5 are the basis for the solution of fixed difference
equations, as discussed in the next section.
To prove Eq. 3.12-3, for example, let m = n + 1, and write
oo oo
Table 3.12-1
Equation
Property Number
ntf(t)]=-Tz-F(z) (3.12-8)
Z yw l_ CF(2)
dz (3.12-9)
T
Z[f(t - kT)U_x(t - kT)] = 2-^(2) (3.12-10)
3V/(()] = F (3.12-11)
a-
n
Z 2 fx(t - kT)f2(kT) = Fl(z)f2(2) (3.12-12)
k=0
1 I F1(w)F2(z/w)
^[/i(0/2(0] = J-. <j>
W
where the contour of integration separates the
poles of Fj(w) from those of F2(z/w) (3.12-13)
lim f(t) = lim F(z), provided that the limit exists (3.12-14)
t—>-0 z—> co
Table 3.12-2
Equation
Property Number
F(s)
F(z) = 2 residues of at the poles of F(s) (3.12-16)
1 — esJz
00
Ink 2tm
F*(s) = - 2 F[s +j = F* s + j (3.12-17)
lc=— oo T T
If F(s) = F1(s)F2*(s), then F(z) = F1(z)F2(z) (3.12-18)
/*( 0 =/(0<(0
where z(0 denotes the impulse train
00
1(0 = 2 (/„(< - nT)
72=— 00
Then
= SF[f(t)i(t)]
which suggests the use of the complex convolution theorem of Eqs. 3.7-7
and 3.7-8. Using the result of Example 3.4-3,
1
/(s) 1 _ e-r*
= 1
Ts = d= j2rrk
2rrk
S = ±j for k = 0, 1, 2, . . .
T
Using Eqs. 3.7-7 and 3.7-8,
1 F(w)
F*(s) = dw
2-rrj Jc 1 - <rr(s-”’>
F(w)
= 2 residues of | _ ^—Ts^wT
at the poles of F(w) (3.12-19)
Replacing els by z,
F(w)
r(*) = 2 residues of — at the poles of F(w)
i - e Tz
Sec. 3.12 Properties of the Z Transform 165
This is identical with Eq. 3.12-16, if the dummy variable w is changed back
to 5. The expression provides a method of obtaining F(z) directly in closed
form, without having to sum an infinite series. Furthermore, in the
analysis of systems, F(s) rather than f(t) may be given. In this case, Eq.
3.12-16 permits the calculation of F(z) without having to find f(t) as an
intermediate step.
Example 3.12-1. Find 2?[te~at].
1
F(z) = residue of at s = —a
(s + a)\ 1 e-'z-1)
d 1 Te~aTz
ds \ 1 — esTz~1 (Z — c-aTy
S——a
The same result follows from Table 3.11-1 and Eq. 3.12-7 or 3.12-8.
Example 3.12-2. Find F(z) corresponding to F(s) = [5(25 + 3)]/[(5 + 1)2(5 + 2)].
F(s)
The residue of at .y = — 1 is
[1 - esTz~l]
5(25 + 3) -T,
-Te-Tz
7sT .y-1
_ds ((5 + 2)(1 — €8rz~1)j_| s_ 1 (z — e~T)2
The residue at s — —2 is
5(25 + 3) 2z
- 2T
.(5 + 1)2(1 - esTz~1) J s=-2 Z — €
Hence
2z Te-Tz
F(z) =
2T
Z — €
—
(z - e-T)2
5(25 + 3)
Z’-1 = 2e-2t - te-
_(s + 1)2(5 + 2)_
and by looking up the Z transform of the two functions of time in the tables.
f F(w) } 1 / .2nk
= ~ ~ E 5 ± J —
\(djdw)[l - €-r,S-"']l„=s±)(2rt/r, T \ T
166 Transform Techniques
Thus
00
. 2nk\
F*0) = - I F\s+j
T k=— oo T7
which is identical with the first half of Eq. 3.12-17. An alternative proof
of this result is suggested in one of the problems at the end of the chapter.
If 5 is replaced by s -f- jilim/T),
p*
. 2?TU 2v-(k + n)^j
s +j
T
The right side may be written
1 00
. 2 tt m \
2 F(s+j = F*(s)
T m=—co T 7
which proves the second half of Eq. 3.12-17.|
Although perhaps the most important application of Eq. 3.12-17 is in
the proof of other theorems, the equation also gives considerable insight
into the sampling of continuous signals. An assumed plot of the frequency
spectrum of a continuous signal f(t) is shown in Fig. 3.12-1#. It is assumed
that the signal is band-limited, so that it contains no frequency com¬
ponents above coc radians per second. The frequency spectrum for F*(s)
is given by
and is shown in parts (b) and (c) of the figure for two different values of
the sampling interval T. The plot consists of the spectrum of the con¬
tinuous signal repeated every 2tt/T radians per second. If 77/T > coc, as
assumed in part (b), the original shape of \F(jco)\ is not destroyed by the
t Equations 3.12-16 and 3.12-17 are given in the form usually used in the literature, but
they are not consistent if fit) has a discontinuity at t = 0. Equation 3.12-16 is correct
if the value of the function at the origin is defined to be f(0-f), while Eq. 3.12-17
assumes that this value is M/(0—) + /"(0 + )]. The inconsistency can be explained by
recalling that Eq. 3.7-8 requires that the integral vanish along DFA and DHA in Fig.
3.7-7, and by carefully examining the integration along these semicircles. Unless
otherwise stated, it is customary to use f(0+) as the function’s value at t — 0. Since
/(0—) = 0 in the one-sided transform, Eq. 3.12-17 should then read
In the proof of Eq. 3.12-17 that is suggested in Problem 3.31, the Fourier series for
the impulse train assumes that /'(/) is an even function of time. The one-sided Laplace
transform then includes the effect of only half of the impulse at the origin, so the extra
term i/(0 + ) is again needed.
Sec. 3.12 Properties of the Z Transform 167
IWujl
I F*(jw)\
\F*(jo,)\
A
■>- co
-3tr/T -tr/T tt IT 3tr/T
Fig. 3.12-1
sampling process. If, however, n/T < coc as in part (c), the original shape
of \F(jaj)\ no longer appears in the frequency spectrum of the sampled
signal. Thus the continuous signal f{t) can be theoretically reproduced
from the sampled signal f*(t) by linear filtering if and only if tt/T > coc.
This statement agrees with Shannon’s sampling theorem.17
The reconstruction of the continuous signal could be accomplished by
an ideal low-pass filter, but approximating an ideal filter requires a large
number of components and results in a great time delay. In many sampled-
data systems, the circuit following the sampling switch does act as a crude
low-pass filter, reducing the high-frequency components and smoothing
out the signal.18
If the 5-plane is divided into horizontal strips as in Fig. 3.12-2, Eq.
3.12-17 indicates that the pole-zero pattern in each strip is the same.
Values of F*(s) at corresponding points in different strips are identical.
Thus F*(s) is uniquely described by its values in the strip —n/T < w <
tt/T.
It is useful to note that the left half of the 5-plane maps into the interior
168 Transform Techniques
co
/V
Fig. 3.12-2
of the unit circle in the 2-plane, as shown in Fig. 3.12-3. If s = a + jco and
2 = peje, where 2 has been written in polar form, the transformation
2 = esT gives
(3.12-20)
Then p = eaT and 6 = coT. For a < 0, a = 0, or a > 0, the magnitude
of 2 is p < 1, p=l (the unit circle), or p > 1, respectively. Note
also that, if co is replaced by co + 2tt/T, the value of 2 is unchanged. Thus
each horizontal strip in Fig. 3.12-2 maps into the entire 2-plane. This is, of
course, consistent with the previous conclusion that the values of F*(s)
for —v/T < co < n/T uniquely determine F*(^) and hence F(z). It also
explains why there are a finite number of poles of F(z), despite an infinite
number of poles of F*(s).
Equation 3.12-18, which is used in the next section, follows directly
from Eq. 3.12-17. If F(s) = F1(s)F2*(s), then
CO
/V
s-plane
Fig. 3.12-3
Sec. 3.12 Properties of the Z Transform 169
Since
f** (* +j^y) =
1 00 / %Tk\
F*(s) = F2*(s) - 2 FAs + j—7 =
1 lc=-oo \ T 1
or
F(2) = ^(2)^(2)
dnf( 0)
dz
22 2z -2
Example 3.12-3. F{z) — . Determinef(nT).
(z - 2)0 - l)2 1 - 4Z-1 + 5z~2 - 2z
By long division,
F(z) = 2z-2 + 8z-3 + 222-4 H-
Thus
m 0
AT) 0
f(2T) 2
/(3 T) 8
/(4 T) 22
This method is often easier than any other when f(nT) is desired for only a few values
of n. A disadvantage is that it does not give a general expression for the «th term in
closed form.
F(z) 2
(z - 2)0 - l)2 z-2 0 ~ If z - 1
z z z
F(z) = 2
z-2 0 - l)2 2 - 1
The reader should verify that, for n = 0, 1, 2,..., this result checks the answer of the
previous example.
2O2 — 2z —1)
Example 3.12-5. F(z) = — ^ — . Find f(nT).
F(z) B C D
+ + 7—:—+
0 -,/)2 0 —j) (2 + y')2 0 + j)
where
1 —F / 1 —/
A =-- , C = -- , B = D = 0
2 2
Sec. 3.12 Properties of the Z Transform 171
Then
F(z) = L±2
2 (~jz - l)2
, i -./
2
m
(jz - l)2
1 -j €~^/2Z
-)--
1 +/ d7Xi2Z
~Y~ (e-^/2z - l)2 2 (e0nt2z - l)2
This example is typical of the manipulations used when F(z) contains complex conju¬
gate poles.
The third method of finding f(nT) is to use the inversion integral, which
may be derived in several ways. For generality, the derivation is per¬
formed in terms of the two-sided transform. From Section 3.7, the
inversion integral for the two-sided Laplace transform is
where the contour C is any path from c — joo to c + j00 within the region
of convergence ac < a < oy . For convenience, the contour is taken as
a straight vertical line. On the basis of the discussion associated with
Figs. 3.12-2 and 3.12-3, it is logical to break this contour up into the
individual sections . . . , —37t/T < co < —rr/T, —tt\T < to < tt/T,
tt\T < co < 377/r, .... Also, since the purpose of the derivation is to
obtain f{nT) from F(z), t is replaced by nT.
1 00 rc+U(2fc+l)jr]/T
f(nT) = -?- T F(s)enTsds (3.12-22)
2tTj *=-oo Jc+Wk-MlT
A
2-plane
>
Fig. 3.12-4
Letting z = eTs and dz = Tz ds, and recalling from Eq. 3.12-20 that the
vertical line a — c (—77/T < 00 <7t/T) corresponds to a circle of radius
ecT in the 2-plane,
Figure 3.12-4 shows two circles of radii px = exp (Toc ) and p2 = exp
(Tac ). The closed contour in Eq. 3.12-23 must be confined to the shaded
area, which is the region of convergence for the direct Z transform.
F(z) is analytic at all points in this region. By the residue theorem of
Eq. 3.6-16,
finT) = ^ [residues of Ffyz71-1] (3.12-24)
at the poles in the region \z\ < px. This equation is perfectly general,
and it can be used for both positive and negative values of n.
Example 3.12-6. F(z) = z[(2 - €)z - 1 ]/[(z - €)(z - l)2], with the region of con-
vergence 1 < \z\ < e. Find f(nT).
zn[(2 - e)z - 1]
F(z)zn-1
"(T- e)(z - iy
For n > 0, the only pole inside the contour of integration is at 2 = 1. The residue at
this pole is
d zn[( 2 — e)z — 1]
n + 1
dz Z — € Jz=l
Hence
finT) = n + 1 for n > 0
Sec. 3.12 Properties of the Z Transform 173
For n = —2, the residue at the pole at 2 = 1 is —1. The residue at the second order
pole at the origin is
- d (2 - e)2 - 1
= 1 + e-2
_dz (2 — e)(z — l)2
=0
SO f( — 2T) = e-2.
— (> F(z)zn 1 dz
2ttJ 3 c
= —2 [residues of F(z)zn~1 at the poles outside contour C] (3.12-25)
In the preceding example, the residue at the pole at 2 = e, the only one
outside of the contour of integration, is
Thus f(nT) = en for n < 0. It is seen that poles of F(z) in the regions
\z\ < pY or \z\ > p2 in Fig. 3.12-4 give rise to components of f(nT) for
n > 0 and n < 0, respectively. In the one-sided Z transform, the dis¬
cussion of Section 3.4 indicates that p2 -> 00, and f(nT) becomes zero for
n < 0. The region of convergence is then the entire 2-plane outside of
the circle of radius pu which encloses all the poles of F(z).
Example 3.12-7. The one-sided transform is F(z) = 22/(2 — 2)(2 — l)2. Find f(nT).
The residue of F(z)zn-X at 2 = 2 is
2zn
= 2(2")
- 1h 2=2
and at 2 = 1 is
d / 227
= -2 (n + 1)
dz\z — 2, J 2=1
so
f(nT) = 2[2n - n - 1] for n > 0
The nature of the terms in f(nT) depends upon the location of the poles
of F(z). Table 3.12-3, showing the effect of pole position in the one-sided
transform, can be constructed from Eq. 3.12-24 or its equivalent. Poles
inside or outside the unit circle yield terms that approach zero or infinity,
respectively, as t approaches infinity. First order poles on the unit circle
(except at z = 1) yield sinusoidal terms of constant amplitude, but higher
order poles produce terms that increase without limit as t approaches
infinity.
Sec. 3.13 Application of the Z Transform 175
When the two-sided transform is used, the form off{nT) depends upon
the region of convergence as well as the location of the poles. If the func¬
tion of time is known to be bounded, the region of convergence must
include the unit circle. In this case, poles inside and outside the unit
circle yield components off (nT) for positive and negative time, respectively.
Any fixed linear system whose input v(t) and output y(t) are defined
only at the discrete instants t = kT can be described by the difference
equation
where the a - s and b/s are constants. By the use of Eqs. 3.12-1 through
3.12-5, the difference equation can be transformed into an algebraic
equation in 2, which can be solved for Y(z). y(kT) can then be found by
the inversion methods of the last section. Note that the transformed
equation will contain y(0) through y(qT — T). These terms represent the
q boundary conditions needed to determine the arbitrary constants in the
classical solution.
Example 3.13-1. Find a general expression for Y(z) when
The constants y{0), t>(0),..., y{2), v{2) are not directly given. Substituting k — — 3, —2,
and —1 into the original difference equation, however, and noting that y{k) and v(k) are
zero for k < 0,
a3y{ 0) = b3v( 0)
a3y{\) + «2?/(0) = M(l) + b2v(0)
a3y( 2) + a2y( 1) + a:y(0) = b3v{ 2) + 62t;(l) + MCO)
m=^ + + +
a3z3 + ^2 2 + axz + a0
2
m
V
Y(z) =
K2"* + "’ + ^z + ^ V(z) (3.13-2)
aaza + • • • + <2,2 + a0
jk for k > 0
v(k)
lo for k < 0
rV
N
hi
22 - 32 + 2
1
(3.13-3)
t H(z) is sometimes called the “pulsed transfer function” or “sampled transfer function”
to distinguish it from H(s).
Sec. 3.13 Application of the Z Transform 177
and can be written down directly from the difference equation, as in Eq.
3.13-2. In Section 2.12, it was shown that H(z)zk is the forced response
to v(kT) = zk. From Eqs. 2.12-37 and 2.12-38,
Equation 3.13-3 enables the output y(nT) to be found by the same three-
step procedure used for the Fourier and Laplace transforms. For a
fl for k = 0
discrete system subjected to the unit delta function v(kT) = ,
V(z) = 1, and |0 for fc ^ 0
d(nT) =S’~1[H(z)] (3.13-5)
where d(nT) is the response to the unit delta function. Equations 3.13-4
and 3.13-5 provide a convenient way of finding a system’s delta response
from its difference equation.
Example 3.13-3. If the discrete system of Fig. 1.9-la is described by the difference
equation of Example 3.13-2, find its delta response.
d(n) = iT-1
2 - 2
This form does not appear in Table 3.11-1, but, by the initial value theorem of Eq.
3.12-14,
2
d(0) = lim - = 0
z—►OO 2 2
By Eq. 3.12-3,
22
dik + 1) =
2-2
where H(s) is the transform of the impulse response h(t). By Eq. 3.12-18,
It is important to note that, while Eq. 3.13-6 can be solved for the output
as a continuous function of time, Eq. 3.13-7 yields the output only at
the sampling instants. The calculation of the output between sampling
instants is deferred until the next section. The determination of H{z)
for more complicated configurations is illustrated by the following
examples.
v
Example 3.13-4. Find H(z) — Y(z)/V(z) for the configurations of Fig. 3.13-1. The
systems are initially at rest, and the sampling switches operate synchronously.
(a)
(b)
Fig. 3.13-1
In part (a) of the figure, Q(z) = Hfz)V(z) and Y(z) = H2(z)Q(z); so H(z) =
Hx{z)H2{z). In part (b), H{z) is the Z transform corresponding to H1(s)H2(s). This
can be written
H(z) = <ar\h(t)]
Hfs)H2(s)
H(z) =2 residues of at the poles of H^H^s)
1 - esTz~1
is often used, but the reader must remember that HyH^z) ^ Hx(z)H2(z).
Example 3.13-5. In the previous example, find y(t) for t — nT if hft) — h2(t) = tcl
for t > 0, T = 1, and v(t) = e'tU-ft).
For the configuration of part (a) of Fig. 3.13-1,
Sec. 3.13 Application of the Z Transform 179
1 d4 1
= e -zn+2 = — (n + 2)(n + l)(n)(n — l)e_” for n > 0
4! dzi Jz=l
which agrees with the answer to Example 1.9-4 at the sampling instants.
For the configuration of part (b) of the figure, h(t) = r3e“76 by Example 1.6-2, so
3
tAe (ze)(z2e2 + 4ze + 1)
H(z) = %
6(2e - l)4
Then
(Zt)2(2262 + 426 + 1)
Y(z) =
6(26 - l)5
and
1 dx
2/00 =
44 dz*
(z” + 3 + 42”+2 + 2”+1)
Jz=l
n2{n + 1):
e~” for n > 0
24
which agrees with the answer to Example 1.9-3.
Example 3.13-6. Figure 3.13-2 shows a common sampled-data system, labeled with
Laplace transformed quantities. Find an expression for F(2).
W(s)
H(s)
Fig. 3.13-2
ETsing Eq. 3.12-18,
Q(z) = V(z) - GH(z)Q(z)
G(z) V(z)
Y(z) = G(z)Q(z) =
1 + GH(z)
The proof of this convolution theorem follows from the basic definition
of the Z transform,
00
U*) = 2 fi(nT)z~n
n= 0
oo
F2(z) = 2 UnT)z-"
n—0
agreeing with Eqs. 1.9-1 and 1.9-2. For the continuous system and
sampling switch of Fig. 3.11-1, Y(z) = V(z)f£[h(t)]; hence the output at
the sampling instants is
n n
y(nT) =2 v(kT)h(nT~ kT) =2 v{nT - kT)h(kT) (3.13-11)
k=0 A:=0
As in Section 3.8, the poles of V(z) and H(z) give rise to the forced (or
particular) and free (or complementary) response, respectively. A system
is unstable if its free response increases without limit as t approaches
infinity. Table 3.12-3 indicates that a system is unstable if and only if
H{z) has poles outside the unit circle, or multiple order poles on the unit
circle. From Eq. 3.13-4, the poles of H(z) are the roots of the characteristic
equation
aQzq + • • • + aiz + aQ = 0
Hence the above conclusions about stability agree with those of Section
2.12.
Sec. 3.13 Application of the Z Transform 181
y(t)
Time-Varying Systems
for a discrete system, where the transformation is with respect to kT, and
where d*(nT, kT) is the response at time nT to a unit delta function
occurring kT seconds earlier. For the configuration of Fig. 3.11-1, the
system function is
H(nT, z) = £T[h*(nT, kT)] (3.13-14)
Then
y(nT) = fT-1[V{z)H(nT, z)] (3.13-15)
The principal difficulty with this scheme is finding H(nT, z). If, however,
the delta or impulse response is known, this can be done by Eq. 3.13-13
or 3.13-14.
182 Transform Techniques
The use of the Z transform in the analysis of Fig. 3.11-1 yields the
output y(t) only at the sampling instants t = nT. Although a number of
methods have been suggested for obtaining the output between sampling
instants, the most useful one seems to be the modified Z transform.
The calculation off£[f{t)] is based upon samples off(t) at the instants
t = nT. In order for the Z transform to contain information about the
function of time at other instants, f{l) should be sampled at t = (n + y)T,
where y is a real, independent parameter which may assume any value
between zero and one. Parts (a) and (b) of Fig. 3.14-1 show an arbitrary
function f(t) and a train of unit impulses i(t). Recall that the product
fit) i(t)
A
oo oo oo oo oo oo
A CO
CO A
/V
1 t
0 T 2T 3T AT
ie)
Fig. 3.14-1
Sec. 3.14 The Modified Z Transform 183
f{t)i(t) gives
oo
/*(<) = 2
n=— oo
f{nT)U,(f - nT) (3.14-1)
In order to sample f(t) at the instants t — (n + y)T, the plot of /'(/) can
be moved forward yT seconds. Alternatively, the plot of f(t) can be
moved backwards yT seconds, as in part (d) of the figure. The product
f(t + yT)i{t) gives
00
f*(t, y)= 2
n—— oo
f(nT+ yT)U0(t - nT) (3.14-2)
The various nomenclatures introduced in Eq. 3.14-4 all stand for the Z
transform of f(t + yT), otherwise known as the modified Z transform
qffm Note that
F(z) = lim F(z,y) (3.14-5)
y^O
Y(z, y) = y) (3.14-8)
where
00
By Eq. 1.9-3, the coefficient of z~n is y(nT + yT), so the infinite series
above is identical with the definition of Y(z, y), thus proving Eq, 3.14-8.
Finally,
y(nT + yT) = £T~1[V(z)H(z, y)] (3.14-10)
Example 3.14-1. Solve Example 3.13-5 for y(t, y). In Fig. 3.13-1, hx(t) = h2(t) = te~l
for t > 0, T — 1, and v(t) = e~tU_1(t).
For the configuration of part (a) of Fig. 3.13-1,
Q{z) = V(z)H1(z)
Y(z, y) = V(z)Hx(z)H2(z, y)
where
Ze
V(z) =
(ze - 1)
ze
Hx{z)
(ze - \y
€ y(gg) ye '(ze)
H2(z, y) = Z[(t + y)e-(*+y>] =
(ze — l)2 ze — 1
Using Eqs. 3.12-7 and 3.12-24,
e y(ze)3 ye y(ze)3
y(n, y) +
(ze - l)5 (ze - l)4
1 J4 y <73
= e~ne~y -(2"+2) + --(zn+2)
4! dzx v 3 \ dz3 V 7 Jz=l
1
= — (« + 2)(« + 1)(«)(« — 1 + 4y)e (n+y) for « > 0
i
References 185
As a check, note that, when y = 0 or 1, y{n + y) reduces to y(n) or y(n + 1), respec¬
tively, in Example 3.13-5. For 0 < t < 1, n = 0 and t = y, while for 1 < t < 2,
n = 1, and t — 1 + y, so
Y(z, y) = V(z)H1H2(z, y)
where
(t + y)3e
HxH2{z, y)=2t
Then
2/(0 = t3 —
for 0 < / < 1
o
Example 3.14-2. Find an expression for Y(z, y) for the configuration of Fig. 3.13-2.
As in Example 3.13-6,
no
(2(0 =
1 + GH{ 0
Then
G(z, y)V(z)
Y(z, y) = Q(z)G(z, y) =
1 + GH(z)
REFERENCES
5. S. Seshu and N. Balabanian, Linear Network Analysis, John Wiley and Sons, New
York, 1959, Section 5.5.
6. Gardner and Barnes, op. cit., p. 346, Eq. 2.1362.
7. R. V. Churchill, Complex Variables and Applications, Second Edition, McGraw-
Hill Book Company, New York, 1960.
8. B. van der Pol, and H. Bremmer, Operational Calculus Based on the Two-Sided
Laplace Integral, Second Edition, Cambridge University Press, Cambridge, 1955,
Section VI. 11.
9. W. R. LePage, Complex Variables and the Laplace Transform for Engineers,
McGraw-Hill Book Company, New York, 1961, Chapter 10.
10. S. Goldman, Transformation Calculus and Electrical Transients, Prentice-Hall,
Englewood Cliffs, N.J., 1949, Section 11.8. v
11. J. A. Aseltine, Transform Method in Linear System Analysis, McGraw-Hill Book
Company, New York, 1958, Chapter 17.
12. L. A. Zadeh, “Frequency Analysis of Variable Networks,” Proc. IRE, Vol. 38, No.
3, March 1950, pp. 291-299.
13. D. Graham, E. J. Brunelle, Jr., W. Johnson, and H. Passmore, III, “Engineering
Analysis Methods for Linear Time Varying Systems,” Report ASD-TDR-62-362,
Flight Control Laboratory, Wright-Patterson Air Force Base, Ohio, January,
1963, pp. 132-150.
14. L. A. Zadeh, “A General Theory of Linear Signal Transmission Systems,” J.
Franklin Inst., Vol. 253, April 1952, pp. 293-312.
15. L. A. Zadeh, “Time-Varying Networks I,” IRE Intern. Conv. Record, Vol. 9, Pt.
4, March 1961, pp. 251-267.
16. Gardner and Barnes, op. cit.. Chapter IX.
17. C. E. Shannon, “Communication in the Presence of Noise,” Proc. IRE, Vol. 37,
No. 1, January 1949, pp. 10-21.
18. J. G. Truxal, Automatic Feedback Control System Synthesis, McGraw-Hill Book
Company, New York, 1955, Section 9.2.
19. W. Kaplan, Advanced Calculus, Addison-Wesley Publishing Company, Reading,
Mass., 1952, p. 569.
20. M. Marden, The Geometry of the Zeros of a Polynomial in a Complex Variable,
American Mathematical Society Mathematical Survey, No. Ill, New York, 1949,
p. 152.
21. Truxal, op. cit.. Section 9.6.
22. J. R. Ragazzini and G. F. Franklin, Samp/ed-Data Control Systems, McGraw-Hill
Book Company, New York, 1958, p. 93.
23. E. I. Jury, Theory and Application of the z-Transform Method, John Wiley and Sons,
New York, 1964, pp. 289-296.
Problems
3.4 By extending the result of Example 3.2-1, find and sketch the complete
output voltage for Fig. 3.2-2#. Assume that the circuit contains no initial
stored energy. Compare the answer with Example 2.12-3.
3.5 If the system function of a fixed linear system is H(s) = s/(s2 + 0.2s +
100), and if the input is the periodic waveform shown in Fig. P3.5, find and
sketch the Fourier series representing the steady-state output.
v(t)
Fig. P3.5
t/wr * = -
.00
|G(a>)|2 dco
H(jw)
! \
1
_
-C0c 0 Wc
Fig. P3.9
3.13 Find the inverse transform of the following functions by partial fraction
expansions, and Tables 3.3-1 and 3.4-1.
1
Fx{s)
(s + 1)V + 4)
F2(s)
(s + 1)V + 4)
F3(s)
0 + 1)V + 4)
252 + F
F*(s)
s(s2 +2^+5)
3.15 Solve Examples 2.6-6 and 2.6-7 and Problem 2.2 by the Laplace transform.
3.16 A system is described by the differential equation
Find the impulse response h(t), and evaluate h(0 +), //(0 +), and /?(0+).
If it is possible, find these initial conditions by the initial value theorem
of Eq. 3.4-15. Check the results by the methods of Chapter 2.
3.17 The circuit in Fig. P3.17 illustrates the application of the Laplace trans¬
form when there is initial stored energy. The circuit is originally operating
18 2 h
vw-w
4
28 18
volts
V/v Wv
2
volts
Fig. P3.17
in the steady state with the switch K open. If the switch closes at t = 0,
shorting out a resistance, find expressions for the currents for t > 0.
Represent the energy stored in the inductance and capacitance at t = 0
by added sources.
3.18 Write the Laurent series for each of the following functions about the
singularity at 5 = — 1. The quantity t should be treated as an independent
parameter.
.st Se st
(a) (b)
S + 1 (s + l)2
Problems 189
Find the residue at 5 = —1 from the series, and also by Eqs. 3.6-17
through 3.6-20.
bmsm + • * • + biS + bo
3.19 If F(s) = show that
ansn + • • • + a^s + ciq
1 r F'(s)
W)ds = z~p
where the contour C encloses the right half of the 5-plane, and where Z
and P denote the number of zeros and poles, respectively, of F(s) in the
right half-plane. Fligher order poles and zeros should be counted
according to their multiplicity. This result leads to the Nyquist stability
criterion.
3.20 If the one-sided transform is F(s) = s/[(s + 2)(j — 1)], find f(t) by the
inversion integral. Repeat for the two-sided transform
Fig. P3.25
dv
y — + v
t -f- 1 dt
find h*(t, r) and H(t, s). Find the step response using transform tech¬
niques, and compare the answer with Problem 1.12.
3.27 Find the system function H(t, s) corresponding to each of the following
differential equations.
3.30 Prove those properties in Table 3.12-1 which are not derived in this
chapter.
3.31 Prove the first half of Eq. 3.12-17 by replacing /(/) by its complex Fourier
series, before transforming the relationship f*(t) = f{t)i{t). Problem 3.3
and Eq. 3.4-8 are helpful.
3.32 If |[/(/)] = z/[(z + 2)(4^2 + 1)], and if/(/) is known to be a bounded
function, find f(nT). Calculate the numerical values of f(nT) for —2 <
n < 4 by the inversion integral and by modifying the partial fraction
method in view of Eq. 3.11-10.
3.33 Find f{nT) for the following one-sided Z transforms, using Eq. 3.12-21,
Eq. 3.12-24, and partial fraction expansions.
z
(a) Fx{z) =
(z + 2)(422 + 1)
2z3 - 2z2 + z
(b) F2(z) =
(z - 1 )(z2 - z + 1)
3.34 Solve Problems 2.14 and 2.15 by the Z transform, again commenting
on the stability of the systems.
3.35 In Fig. 3.13-2, G(s) = 1 /(s + 1), H(s) = 1, and the ideal sampling
switch has a period T given by e~T =
(a) Find the step response r(nT) by the Z transform.
(b) Find the step response r[(n + y)T] by the modified Z transform.
(c) Determine the difference equation which relates v(t) and y(t) at the
sampling instants.
Problems 191
3.36 In Fig. 3.13-2, G{s) = [K(s + 1.04)]/[s(s + 0.692)], H(s) = 1, and the
period of the sampling switch is T = 1.
(a) By a root locus diagram, determine the values of the real constant K
for which the system is stable.
Cb) Let v(t) = 0, but assume that some disturbance produces the signal
^(0) = q{\) = 1. For K = 1, find y{nT).
3.37 A sampling switch is added at the right of H{s) in Fig. 3.13-2, and it
operates in synchronism with the original one. Find expressions for Y(z)
and Y(z, y).
3.38 A “sampler with a zero order hold” is described by the output-input
relationship y(nT + yT) = v{nT) for integral values of n and for 0 <
y < 1. It samples the input every T seconds and maintains the last
sampled value between sampling instants, as in Fig. 1.8-lc.
v*(t) y(t)
hi (t)
(a)
v(t) ho(t) =
y(t)
Sampler with —>
zero order hold a-t-'W-M
(b)
Fig. P3.38
4.1 INTRODUCTION
and the associated topic of functions of a matrix are the most important
topics in this chapter.
011 a\2
@rnl d m2 0mn
then it will be shown that the set of linear equations can be written as
Ax = y by a suitable definition of the “product Ax.” Certainly, this
expression is considerably simpler in form than the set of linear equations.
This is one of the major reasons for using matrices. A matrix equation,
or a set of matrix equations, contains a great deal of information in a
compact form. Without the use of this compact notation, the task of
analyzing sets of linear equations is quite cumbersome.
Consider then, the rectangular array of ordered elements of Eq. 4.2-1.
The typical element in the array may be a real or complex number, or a
function of specified variables. A matrix is a rectangular array such as
shown, but distinguished from simply a rectangular array by the fact
that matrices obey certain rules of addition, subtraction, multiplication,
and equality. The elements of the matrix an, al2, . . . , aiS are written
with a double subscript notation. The first subscript indicates the row
where the element appears in the array, and the second subscript indicates
the column. A matrix is denoted here by a boldface letter A, B, a, b, etc.,
or by writing the general element [ai}\ enclosed by square brackets. The
columns of the matrix are called column vectors, and the rows of the matrix
are called row vectors. A matrix with m rows and n columns is called an
194 Matrices and Linear Spaces
<*i
Cl2
a =
a ii All elements
zero
a 22
Diagonal matrix =
All elements
zero a nn
0 0 0 • • • 1
(e) Null matrix. A matrix which has all of its elements identically equal
to zero is called a zero or null matrix.
(/) Transpose matrix. Transposing a matrix A is the operation whereby
the rows and columns are interchanged. The transpose matrix is denoted
Sec. 4.2 Basic Concepts 195
by Ar. Thus, if A = [aij], then AT = [aH], i.e., the element of the ith
row,y'th column of A appears in theyth row, ith column of A2’. If A is an
(m x n) matrix, then A2 is an (n x m) matrix.
(a) Symmetric matrix. A square matrix, all of whose elements are real,
is said to be symmetric if it is equal to its transpose, i.e., if
A = AT or aij = aH (i,j=\,...,n)
This, of course, implies that the elements on the principal diagonal are
identically zero.
(c) Conjugate matrix. If the elements of the matrix A are complex
0ciij = a+ jfiij), then the conjugate matrix B has elements bij =
a— jpiy This is written in the form B = A*.
(d) Associate matrix. The associate matrix of A is the transposed
conjugate of A, i.e., associate of A = (A*)21.
(e) Real matrix. If A = A*, then A is a real matrix.
(/) Imaginary matrix. If A = —A*, then A is pure imaginary.
(g) Hermitian matrix. If a matrix is equal to its associate matrix, the
matrix is said to be Hermitian, i.e., if A = (A*)2", then A is Hermitian.
(h) Skew-Hermitian matrix. If A = —(A*)2", then A is skew-Hermitian.
Elementary Operations
2 "4 2"
-3 0 1
A + B = B + A Commutative
A + (B + C) = (A + B) + C Associative
196 Matrices and Linear Spaces
“ 2 4" “2 -2“ “0 6“
3 1 _3 0_ -6 1_
A = B (4.2-4)
if an only if a= bij.
Vi = allx1 + a12x2
(4.2-5)
V'2 — + * 22X2
y = Ax (4.2-6)
where
Vi a\2 xx
y = , A = and x =
_9*21 a22_ _X2
Ax = (4.2-8)
xi = 1 bjkzm j = 1.2
k=1
Vi =1aa1
3=1 k=1
i = 1 2 , (4.2-10)
Since the order in which the summations are taken can be interchanged,
2/2
Vi
k
2=1 ( 'zL jk 'k>
i = 1, 2 (4.2-11)
y = ABz (4.2-12)
C = AB (4.2-13)
or
r 2
2
_3 = 1
jk
C = AB
_k=
2 1^ik^kj (4.2-14)
Thus the i,jth element of C is the sum of the products of the elements of the
zth row of A and the corresponding elements of theyth column of B. The
resulting matrix C is (m x p).
If the number of columns of A is equal to the number of rows of B, the
two matrices are said to be conformable, in that the product AB exists.
For the case where A is (m x n) and B is (n x m), then both products AB
198 Matrices and Linear Spaces
(a)
"1 r "1 2" "i r
AB
_1
VO
<N
_2 2 0 1
i
(b)
"1 2“ ~i -r '5 3"
BA ^ AB
0 1 2 2 2 2
(c)
"i r '-i r "0 0”
_1
o
o
2 2 i -l
i
Note that AB = [0] does not imply that A = [0] or B = [0].
Premultiplication or postmultiplication of a
Scalar Multiplication.
matrix by a scalar multiplier k multiplies each element of the matrix by k.
The typical element of the product kA is katj.
Example 4.2-4. Perform the indicated multiplication of the matrix by the scalar 2.
_1
"1 “2 —2~
1
_2 3_
i
n i = , m
'zo = 2 aikbjcj
k=1 i 15 • • • i p
Sec. 4.2 Basic Concepts 199
^ij ji ®kj
k=l k=1
IA = AI = A
Cl 2 od2 2_
to
to
_#21 #22_ 11
1
~dx 1 0 -
#11 #12 &iid\\ ^12^11
Premultiplication:
0 d22 #21 #22 ^21^22 ci 2 2d22
aw a\2 ai3
'A! A3
A =
h>l a22 a23
_A2 a4_
_a3\ a32 a33_
where
a.21 a 22 a23
A-i — [/hi] A2 — A3 — [^12 ^13] A4 —
A + B =
_A + B A + b 2 2 4 4
AB =
_(A B + A B2) (A B + A B4)_
2 1 4 2 3 4
'0 2 ■ 1 0 3"
AB = 31
1 -1 °J -1 2 1
0 1 -1_
—
0 “ 1 0 3" ”3"
+ [0 1 -1]
1 -1 -1 2 1 0
2 4 2~ "0 3 -3“
_
~—2 7 -r
+
2 -2 2_ _0 0 0 2 -2 2
dt dt dt
Sec. 4.2 Basic Concepts 201
From this definition it is evident that the derivative of the sum of two
matrices is the sum of the derivatives of the matrices, or
^22(0 dt
CM
rH
...
A(t) dt = (4.2-21)
>
) dt (4.2-22)
Example 4.2-7.
~t r
Find Qq(A), if A
_i t_
Q$(A)
't
t dt t
Jo
202 Matrices and Linear Spaces
4.3 DETERMINANTS
The theory of determinants is also useful when dealing with the solution
of simultaneous linear algebraic equations. Determinant notation simpli¬
fies the solution of these equations, reducing the solution to a set of rules
of procedure. As an example, the set of equations
awx\ ~b ^12*^2 — yi
(4.3-1)
*21*^1 “b Cl22X2 ~ y%
can be solved by finding an expression for x1 from the first equation, and
then substituting this expression into the second equation. The result of
performing this operation is the solution
This solution assumes that the denominator (tfntf22 — a12a21) is not zero,
Equation 4.3-1 can be written in matrix notation as
Ax = y (4.3-2)
where
~*n *i2~ xx ~y{
A = , X = and y =
_ci 21 a 22_ _x2_ _y 2_
*11 *12
IAI =
*21 a22
a 11 y1
= (Ml - 2/l«2l)
*21 2/2
Sec. 4.3 Determinants 203
are formed, then the solution to Eq. 4.3-1 can be expressed in terms of
these determinants as
lAii
*i =
IA |
provided |A| ^ 0.
The definition of the value of the determinant of any square matrix A
follows. The determinant of the (n x n) square matrix A, written as |A|,
has a value which is the algebraic sum of all possible products of n elements
which contain one and only one element from each row and column,
where each product is either positive or negative depending upon the
following rule: Arrange the possible products in ascending order with
respect to the first subscript, e.g., ar3a22a31 • • • . Define an inversion as the
occurrence of a greater integer before a smaller one. The sign of the
product is positive if the number of inversions of the second subscript
is even; otherwise it is negative. For example, the sequence 321 has three
inversions: 3 before 2, 3 before 1, and 2 before 1.
I A| # 2i #22 #23
a 1
3 a 32 a 33
«n02 2^33 0 4-
# 11#2 3^32 1 -
# 10 # 2 1^33 1 -
^12^23^31 2 +
«13«2 1^32 2 +
#13^2 2^31 3 -
2. The value of the determinant is zero if all the elements of any row
(or column) are zero, or if the corresponding elements of any two rows
(or two columns) are equal or have a common ratio.
3. The value of the determinant is unchanged if the rows and columns
are interchanged.
4. The sign of the determinant is reversed if any two rows (or two
columns) are interchanged.
5. The value of the determinant is multiplied by a constant k if all the
elements of any row (or column) are multiplied by k.
6. The value of the determinant is unchanged if k times the elements of
any row (or column) are added to the corresponding elements of another
row (or column).
Minors. If the zth row and y'th column of the determinant |A| are
deleted, the remaining n — 1 rows and n — 1 columns form a determinant
|M^|. This determinant is called the minor of element aij. For example,
if the determinant A is given by
#n #12 #13
#31 #33
#21 #23
Cl2 ~ (—1)
#31 #33
Sec. 4.3 Determinants 205
Example 4.3-2. Evaluate the minors and cofactors of the third order determinant
shown in Example 4.3-1.
#2 2 #23
|M„| = — <722033 #23#32 — Cll
#32 #33
#21 #23
|M„| = — #21#33 #23#31 — — C12
#31 #33
#21 #22
|M13| = L— #21#32 #22#31 — E^13
#31 #32
#12 #13
|m21| = — #12#33 #13#32 — — C'21
#32 #33
#11 #13
|M„| = #11#33 #13#31 — C22
#31 #33
#11 #12
|m23! = — #11#32 #12#31 — C23
#31 #32
#12 #13
|m31| = — #12#23 #13#22 — T31
#22 #23
#11 #13
|m32| = — #11#23 #13#21 — C32
#21 #23
#11 #12
|m33| = — #11#22 #12#21 — T33
#21 #22
If the results of Examples 4.3-1 and 4.3-2 are compared, it is seen that
the determinant of the third order matrix A can be expressed in terms of the
elements of a single row or column and their respective cofactors. Thus
<
c„ + +
S3
II
products of the elements of any single row or column and their respective
cofactors. Thus
n
ro -1 H
A = 1 2 0
2 0 2
|M„|=4 Cu = 4
| Af ^21 = 2 U12 = —2
\M13\ = 4 C18 = 4
|M12|=2 C12 = —2
|Af22| = -2 C22 = -2
\Mn\ = -l C32 = 1
Using the six fundamental properties of a determinant, the value of the determinant
can be found by reducing the determinant to a diagonal determinant. The value of the
determinant is then the product of the diagonal elements. The following steps show
how this method is applied.
Step 1. Multiply the third row by ( — -J-) and add this to the first row.
— 1 -1 0
I A| 1 2 0
2 0 2
-1 -1 0
I A| = 0 1 0
2 0 2
Sec. 4.3 Determinants 207
-1 -1 0
|A| 0 1 0
0 0 2
-10 0
I A| 0 1 0
0 0 2
The value of the determinant is then
(4.3-5)
(elements in the /th column replaced
2 a3kCji
j=1 ~ 0 k 1
by elements in the kih column)
Using the Kronecker delta notation,
10 i^k
^ik
1 i = k
Eqs. 4.3-4 and 4.3-5 can be combined into the more useful relationship
Pivotal Condensation1 -
in the same column as aij. The elements aqk, aqj, aik, and aij are then
used to form a second order determinant, with the elements kept in their
proper order. Form all such second order determinants with the pivot
term as one of the elements. The original determinant can now be ex¬
pressed as an n — 1 order determinant using the second order determinants
as elements, and 1 /a^r2 as a multiplying factor. The position of the new
elements can be found by subtracting one from each of the subscripts
of the element in the second order determinant that lies on a diagonal
with the pivot term, if this element lies below the pivot term; the sub¬
scripts are unchanged if this diagonal element lies above the pivot term.
By repeating this procedure, the value of va determinant of high order
can be computed by successively reducing the order of the determinant
by one.
Example 4.3-4. Write the given determinant as a determinant of second order deter¬
minants by means of pivotal condensation.
1 3 1 4 1 2
4 5 4 6 4 1
-7 -10 -7
1 3 1 4 1 2
I A| -4 -4 3
3 5 3 8 3 9
-6 -14 -3
1 3 1 4 1 2
4 6 4 2 4 5
Again using the new anas the pivot term, after multiplying the first and second columns
by -1,
7 10 7 -7
4 4 4 3 -12 49
- -302
7 10 7 -7 38 21
6 14 6 -3
Note that each step required only the computation of a set of second order determinants.
Sec. 4.3 Determinants 209
Multiply all the rows of the determinant, except the row containing the
pivot term, by the value of the pivot term. This multiplies the determinant
by the pivot term raised to the n — 1 power. To keep the value of the
determinant unchanged, the resulting determinant is divided by the pivot
term raised to the n — 1 power. Using a±1 as the pivot term, the result of
performing this step is
a ii a 12 a 13
1
|A| = #21#11 #22#11 #23#11
a ii
#31#11 #32#11 #33#11
The first row is then multiplied by the term directly beneath the pivot
term (in the original determinant) and is then subtracted from the row
containing that term.
This process is repeated until all the terms in the column of the pivot
term are zero except for the pivot term, i.e.,
Product of Determinants
Derivative of a Determinant
= C ■ (4.3-8)
oaa dau \*=i
an 0*0 a12(x)
|A| = = «u (x)a22(z) — a2l{x)ai2{x)
tf2i0*0 fl22(r)
Sec. 4.4 The Adjoint and Inverse Matrices 211
If A is a square matrix and CiS is the cofactor of atj, then the matrix
formed by the cofactors CH is defined as the adjoint matrix of A, i.e.,
Thus the adjoint matrix is the transpose of the matrix formed by replacing
the elements aij by their cofactors.
Example 4.4-1. Find the adjoint matrix of
'1 0
A = 2 3
1 2
Adj A = [C„] =
Therefore the product of the matrix [a^] and the adjoint matrix [CH] is
equal to
[««][C„] = IA | I (4.4-3)
which is a diagonal matrix with all its elements equal to the determinant
of the coefficient matrix A. If both sides of Eq. 4.4-3 are divided by |A|,
212 Matrices and Linear Spaces
provided |A| ^ 0,
ama = i (4.4-4)
|A|
From this equation it seems natural to define the matrix (Adj A)/|A| as
the inverse or reciprocal of A, such that
AA-1 = I (4.4-5)
where
It is evident that A_1A = I, i.e., that a matrix and its inverse commute.
Only square matrices can possess inverses.
Example 4.4-2. Find the inverse of A and verify that AA_1 = I, if
1_
1
(N
co
2 3 Adj A
i
A = , A - 1, A-1- , -
1 2 | A| -1 2
"2 2“ " 1 r
0 1 2
1 _ B 1 - -
_1
r-H
o
(N
1
1
B-1A_1i_
=
i
i
3 2
(4.4-8)
dt
This relationship can be derived by considering
7 [A-1(0A(0] = 7 = [0]
dt dt
This is
y [A_1(0] = — A_1(<)A(0A-1(0
dt
Note that this is not the same as {dAjdtY1.
0 1'
A (/) =
-1 -/
The inverse of A is
—t -r
A ~\t) =
1 o
Clearly,
-1 O'
T [A~\t)] =
dt o o
Checking this result with Eq. 4.4-8,
1
1_
1
7
o
o
“-1 0"
7
■♦•a
1
i
— [A ~\t)} = -
i—
1_
1_
1_
_1
o
o
o
o
o
dt
i
1
1
214 Matrices and Linear Spaces
The rows and columns of a matrix are called row vectors and column
vectors, respectively, in Section 4.2. This is simply an extension of the
more familiar concept of vectors in two- or three-dimensional spaces
to an ^-dimensional space. When n is greater than three, the geometrical
visualization becomes obscure, but the terminology associated with the
familiar coordinate systems is still quite useful. For example, the coordi¬
nate system having the unit vectors
-o~
~r -(T
0
0 l
0
0 0
0
0 5 0
> *2 — J • • • > Ifl
• • •
• •
0
0 0
1
Scalar Products
The scalar product (or inner product) of two vectors x and y is written
as (x, y) and is defined as
both x and y are real, Eq. 4.5-1 reduces to the more familiar form
n
Y
.A. -
-
"1 + f
y —
\7 -
"i + r
2 -V -
The scalar product (x*)ry is equal to <x, y> = (1 — j)(1 + j) + 2(2j) = 2 + 4y.
This is not equal to, but is the conjugate of, x^y* = (1 + /)(! — j) + 2(—2j) — 2— 4j.
Outer Product
• • • '1/ ^
XnVl* XnV* ^no m
Orthogonal Vectors
Length of a Vector
Unit Vectors
X = (4.5-8)
X, x>
Example 4.5-2. Find the unit vector x corresponding to the vector x, where
1 +j 2
.2
v
The scalar product <x, x) = 10. Therefore
~1 +,/2 "
A
VIo
X =
2-7
V io
Linear Independence
a1b1
@2^2
^3^3
AB
where
A = B =
Assume that A has m zero elements and B has q zero elements, where
q < m. If the zero elements of A do not occur where the zero elements of
B are located, then the total number of zeros in the product AB is m + q.
If the zero elements of B are located where the zero elements of A occur,
then the total number of zero elements in the product AB is m.
Returning to Eq. 4.5-9 and linear independence, the latter can be
expressed in terms of the rank of the matrix formed by the elements of the
m vectors xl5 x2, . . . , xw. This matrix is
^11 X12 x1 m
X21 X22 X 2m
A = m < n
Xn 1 Xni
X nm
If the rank of the matrix A associated with these m vectors is less than
m, i.e., r < m, then there are only r vectors of the set which are linearly
independent. The remaining m — r vectors can be expressed as a linear
combination of these r vectors. Therefore a necessary and sufficient
condition for the vectors to be linearly independent is that the rank of A
be equal to m.
Gramian
(a) the sum of any two vectors in S is also a vector in the set;
(b) every scalar multiple of a vector in the set is a vector in the set; and
(c) the rules for forming sums of vectors, and products of vectors with
scalars have the following properties:
The most common example of a linear vector space is the set of all vectors
contained within a three-dimensional Euclidian space. For example, all
the forces acting on a space vehicle constitute a vector space, where the
forces are described as vectors in the particular coordinate system chosen.
Sec. 4.5 Vectors and Linear Vector Spaces 219
If a set of vectors xl5 x2, . . . , xm are contained in the space S, then the
set of all vectors y which are linear combinations of these vectors, i.e.,
y = Mi + ^2*2 + ‘ (4.5-12)
T " 0 " T
Xi = 1 x2 = 1 x3 = 2
_0 _ _1 _ _1 _
T T ~r
Xl = 1 , x2 = 2 , x3 = 3
_1_ _3_ _2_
specified in terms of the orthogonal basis (1,0,0), (0,1,0), (0,0,1) span a three-
dimensional space. Specify any vector y of this space in terms of the vectors xx, x2, x3,
and in terms of the basis vectors.
Since the Gramian of these three vectors is
3 6 6
G = 6 14 13 5*0
6 13 14
the three vectors are linearly independent, and therefore they span the three-dimensional
space. Thus any vector y in the three-dimensional space can be expressed as a linear
combination of these three vectors as y = A^Xj + k2x2 + k3x3.
If the vectors x1? x2, and x3 are used as a basis for the three-dimensional space, then,
relative to this basis, the vector y is
Fig. 4.5-1
Sec. 4.5 Vectors and Linear Vector Spaces 221
Relative to the particular orthogonal system of coordinates (1,0, 0), (0, 1,0), and
(0, 0, 1), the vector y is
(k i + A'2 + k3)
y = (k1 + 2k2 + 3A:3)
k = (yi, x2)
<yi» yi)
Note that k — 0 if and only if (y1? x2) = 0, i.e., x2 is orthogonal to yl5
in which case take y2 = x2. Generally,
<yi» x2)
y2 = x2
<yi, yi) Yl
The geometrical idea behind the process is that any vector, in particular x2,
may be decomposed into two components, one of which is parallel to yx
and the other perpendicular to yx. To obtain the latter, which is called
y2 here, the component of x2 in the direction of yx is subtracted from x2.
In a similar fashion, y3 is written as y3 = x3 — k2y2 — ^yi, i.e., as the
component of x3 perpendicular to the plane defined by yx and y2. Ana¬
lytically, if the vector y3 is to be orthogonal to both y2 and yx, this leads to
the equations
(yi, x3) = k2(y,, y2> + ylt y,> = kx(yx, y,>
3—1
y<>-v
y> = xs - 2 y, O' = 2, 3m)
(=i <y<, y<>
The y/s now form an orthogonal set, and the unit vectors
y< (4.5-13)
ft =
T ~r _-r "-r
l
y-2 = 2 - 2 i = 0 and y2 = —= 0
V2
_3_ _i_ i_ i_
Similarly,
T ~-r T r sin
y3 = 3 — isi 0 - 2 1 = 1
2 i 1 21
and
-r
A
y3 2
V6
-1
The orthonormal basis is then
T ~-r ~-r
1 A
i i
1 0 2
’ y!-V2
Yl_ V3
LiJ lj
’ y3 “ V6
L-iJ
Reciprocal Basis
The solution for the k/s of Eq. 4.5-12 can be simplified by defining a
set of vectors r,- such that
Given a basis for a space, it is not difficult to show that there always exists
a set of vectors such that a set of relations of the form of Eq. 4.5-14 is
Sec. 4.6 Solutions of Linear Equations 223
satisfied. The vectors r1} r2, . . . , rm are linearly independent and thus
span the ra-dimensional space spanned by the basis vectors xt. Therefore
they constitute a basis for the space. Owing to the relationship (Eq.
4.5-14) between this basis and the basis formed by the xt- vectors, the basis
formed by the ri vectors is called a reciprocal basis.
The principal use of the reciprocal basis is in finding the constants ki
of Eq. 4.5-12. If the scalar products of both sides of Eq. 4.5-12 are taken
with rl9 the result is
Cramer’s Rule
is given. The a{j and yi are a known set of numbers. The n unknowns xi
are to be determined. Note that the number of equations is equal to the
number of unknowns. By the use of determinants or matrices, a systematic
procedure exists by which the unknowns generally can be determined.
The set of equations, Eq. 4.6-1, can be written in the compact forms,
n
2 aijxj = yit /=1,2, ...,n or Ax = y (4.6-2)
3= 1
224 Matrices and Linear Spaces
where
*1 Vi
x2 V2
• •
X = and y —
• •
• •
_y n_
By multiplying both sides of Eq. 4.6-2 by the cofactor Cik and summing
over all values of i between 1 and n, it follows that
n n n
2 ^ikVi
i=l
Xk = k — 1, 2, . . . , n or x = — y = A xy (4.6-5)
|A| |A|
The numerator of the expression for xk is simply the determinant of A
with the /cth column replaced by the column formed by the right side of
Eq. 4.6-1. Thus Cramer’s rule for a solution by determinants can be
stated as follows:
For a set of n linear algebraic equations in n unknowns aq, x2, . . . , xn,
a solution for the unknowns exists if A is a nonsingular matrix. The value of
a given variable, xk, for example, is the quotient of two determinants. The
denominator of the quotient is the determinant of the coefficient matrix.
The numerator of the quotient is the determinant of the coefficient matrix
with the A:th column replaced by the column consisting of the right-hand
members of the set of n equations.
Example 4.6-1. Determine the solution for the two simultaneous equations
a?i 4- 2x2 = 0
xx — lx2 = 2
using Cramer’s rule.
Sec. 4.6 Solutions of Linear Equations 225
"1
A =
1
The determinant of A is equal to —4. Replacing the first column of A by the right-hand
members of the simultaneous equations, the first unknown x1 is
0 2
2 -2
*i = = 1
-4
1
_1
o
" r
H*
_1
_x2_ _2_ i
L $J
1
i
This solution, which agrees with that determined by Cramer’s rule, can be verified by
inspection.
If the right-hand members of Eq. 4.6-1 are zero, the set of equations is
said to be homogeneous. In this case, the numerator of Eq. 4.6-5 vanishes.
Consequently, if the determinant |A| does not vanish, the set of equations
has only the trivial solution x = 0, i.e., aq = 0, x2 = 0, . . . , xn = 0. If
the determinant |A| does vanish, two or more rows or columns of A are
linearly related. Then a ^-parameter family of solutions can be obtained,
where q is the degeneracy of A.
Assume that the rank of the coefficient matrix is r. The values of r
variables can then be expressed in terms of the other q = n — r variables
by the following procedure:
2. Form r equations with the r unknowns on the left side of the equation,
and the remaining q — n — r unknowns on the right side.
3. Solve for the r unknowns in terms of the q = n — r unknowns by the
use of Cramer’s rule.
<bn, x) = 0
If x is to satisfy Eq. 4.6-6, and hence Ax = 0, it must be simultaneously
orthogonal to all the vectors b1? b2, . . . , bn. But if the rank of A is n,
then the b/s are linearly independent and utilize all the available n dimen¬
sions. No vector can be found which is orthogonal to all the b’s. However,
if A is of rank n — 1, there is one linear dependency relationship between
the b’s. Thus one of the n dimensions of the linear vector space is not
occupied and is available for x. Similarly, if the rank of A is n — 2, two
dimensions are available. Hence there are two linearly independent vectors
which satisfy Eq. 4.6-6. In general, if the degeneracy of A is q, so that the
rank of A is n — q, then q dimensions of the linear vector space are avail¬
able for x. The number of linearly independent solutions of the homo¬
geneous system of linear equations is equal to q.
The preceding discussion was directed toward the case in which the
number of equations and number of unknowns are equal. If the number
of unknowns, n, exceeds the number of equations, m, the rank of A is less
than n. Nontrivial solutions always exist for this case. If m > n, non¬
trivial solutions exist only if the rank of A is less than n. In both cases,
the number of linearly independent solutions is n — r, where r, assumed
less than n, is the rank of A. These solutions can be found by the three
steps indicated above.
Sec. 4.6 Solutions of Linear Equations 227
The rank of the coefficient matrix is two, i.e., the highest order array having a non¬
vanishing determinant is obtained by omitting the first and third columns and the
second row. Consequently, a set of linearly independent equations is
2x2 + 3x4 = —4xx — x3
x2 + x4 — —2xx
-1 5
and x2' = i
1 3
-1 —3
Normalization of x/by Hx/H = 2 and of x2'by ||x2'|| = Vl 1/2 yields the orthonormal
solutions
f ’-l"
-1 1 5
and x./ — .—
1 2vTl 3
-1 -3
The fact that xl5 x2, x/, x2', x/ and x2' all are solutions to the original equations can be
determined by substitution.
The rank of the coefficient matrix is equal to 3. Therefore the unknowns are pro¬
portional to the cofactors of any row of the coefficient matrix. Calculating C4j,
-2 1 -1
Cn — 1 -3 0 = -2
1 0 1
1 1 -1
C 42 1 -3 0 = -10
2 0 1
1 -2 -1
C43 — 1 1 0 -4
2 1 1
1 -2 1
C 44 — 1 1 -3 = 14
2 1 0
n
t Compare Eq. 4.3-6, ^ akjCij = |A|<5,-*., and Eq. 4.6-2, aijxj = yiy under the con-
j=i i=1
ditions that |A| = y,- = 0, / = 1, 2,. . ., n.
Sec. 4.6 Solutions of Linear Equations 229
These equations comprise one independent solution and all scalar multiples of it.
The general procedure for the homogeneous case can be utilized for the
nonhomogeneous case in which the number of equations is less than the
number of unknowns. Such a case is
Equation 4.6-9 may be solved using the procedure outlined for homo¬
geneous systems. If xn+1 is then set equal to —1, the solution to Eq.
4.6-8 can be obtained. If it turns out that Eq. 4.6-9 has a solution only
when xn+1 — 0, then Eq. 4.6-8 has no solution. Consequently, Eq. 4.6-8
has a solution only when the coefficient matrix of Eq. 4.6-8 and the
coefficient matrix of Eq. 4.6-9, called the augmented matrix, have equal
rank.
x1 + 3x2 + x3 — 2x4 = 4
2xx + 5^2 — 2x3 — 5x4 = 10
The coefficient matrix and the augmented matrix are of rank 3; therefore the set of
equations possesses a solution. The square array formed by eliminating the third
column has a nonvanishing determinant. The set of equations can be written as
x4 4- 3x2 — 2x4 = 4 — x3
2xx -f 5x2 — 5x4 = 10 4- 2x3
230 Matrices and Linear Spaces
2 (10 + 2x3) -5
and
12 (6 + 3x3)
x4 = 1 3 (4-x3) = 0
2 5 (10 + 2x3)
Since x3 is arbitrary, the complete solution is then
xx = 10 -f- 11c
x2 = —2 — 4c
x3 = c
xA 0
where c is an arbitrary constant.
an Vi
x9 -x, H-
a12 aX2
a21 V'2
X2 -xx H-
a22 a22
Thus the slopes of the two lines of Fig. 4.6-1 are equal, i.e., the lines are parallel. Conse¬
quently the lines do not intersect, and no solution vector exists.
If | A| 7^ 0, the lines intersect at one and only'one point, given by the solution
x2
(b) Homogeneous. Again two possibilities exist. Since yx = y2 = 0, both lines pass
through the origin. If the slopes are unequal, i.e., if |A| ^ 0, then there is one solution.
It is the trivial one, xx = x2 = 0.
If |A| =0 and q = 1, then both lines have the same slope. Thus the lines coalesce,
and any point on the lines is a solution. Since the equation of the coalesced lines is
a 11x1 + a12x2 = 0 = a2ixx + o22x2, the one independent solution is given by
x2 aii a2i
xx aX2 @22
In terms of the preceding discussion, n = 2, but r = 1. Hence omit one of the original
equations (either one), and the solution is
a 9i
Xo —-x1
a22
which are equal.
Note that
a ii @21
bx and b2
@12 @22
They are linearly related and hence parallel in the plane. The solution x is orthogonal
to both bi and b2.
232 Matrices and Linear Spaces
(c) Dependent nonhomogeneous. The equation is axxxx + aX2x2 = yx. This can be
written in homogeneous form as axxxx + aX2x2 + x3yx = 0. In order to solve as a
homogeneous equation, the form a12x2 = — axxxx — x3yx is written. The solution is
aw V\
x2 —-xx — x3 —
a 12 tx12
This is the equation of a plane in three dimensions, and any point on the plane is a
solution. In particular, along the line in the plane for which x3 = —1, the solution is
an . Vi
x2 --xx H-
a i2 &12
Any point on this line is a solution. For this simple case, of course, this equation is
simply the original equation in rewritten form.
y = Ax (4.7-1)
f The terms eigenvalues and eigenvectors are frequently used in place of characteristic
values and characteristic vectors, respectively. The terms latent roots and latent vectors
are also used.
Sec. 4.7 Characteristic Values and Characteristic Vectors 233
121 - A| = 0 (4.7-4)
The nth. order polynomial in 2, given by Eq. 4.7-4, is called the character¬
istic equation corresponding to the matrix A. The general form for the
equation is
m =a- — 4> - • a — aj
and 2 is again set equal to zero, it follows that
#1 = (^1 + ^2 + * * * + An)
Thus the sum of the diagonal elements of a square matrix is equal to the
sum of its characteristic values. Because of its importance, the sum of the
diagonal elements of a matrix is given a title, namely, trace of the matrix.
Hence the above can be written as
#i = T1
#2 == \{a1T1 + T2)
as = ~i(^2^i + a\T2 + T3) ^ 7-11)
ax = -Tx = -(2 + 1 - 1) = -2
'5 3 r
A2 = 4 2 3
4-2 7
T4 -4 17
A3 = 13 3 11
13 11 3
Modal Matrix
-jt f|AI — A|
d/1 '
indicating that 2, is a repeated root of Eq. 4.7-4. But this has been ruled
out by the assumption that the characteristic values are distinct. Thus
[2,1 — A] is of rank n — 1, and application of Eqs. 4.6-7 and 4.4-1 then
shows that the columns of the modal matrix are proportional to any
t The term “mode” is used because, as later chapters show, the modes of dynamic
behavior of a linear system can be expressed in terms of motion along the characteristic
vectors.
236 Matrices and Linear Spaces
nonzero column of Adj [XJl — A]. Since the columns of Adj [At-I — A]
are linearly related for a given X^ each choice of Xt specifies only one
column of the modal matrix.
Example 4.7-2. Find the characteristic values and a modal matrix corresponding to
the matrix A, where
~2 -2 3“
A = 1 1 1
1 3 -1
A — 2 2 -3
-1 A — 1 -1 = 0 = A3 — 2A2 - 5A f 6 = (A - 1)(A + 2)(A - 3)
—1 —3 A T 1
"o ii -ir
0 1 -1
0 -14 14
For A = 3, the adjoint matrix is
'5 1 4“
5 1 4
5 1 4
Since the characteristic vectors are uniquely determined only in direction, these vectors
can be multiplied by any scalar and still satisfy Eq. 4.7-12. Consequently, a modal
matrix is
"-i ii r
M = l l l
1 -14 1
Each column of the modal matrix is a characteristic vector which spans a one¬
dimensional vector space. The three columns of the modal matrix form a basis in the
corresponding three-dimensional space.
Sec. 4.7 Characteristic Values and Characteristic Vectors 237
The preceding discussion considered the modal matrix when the char¬
acteristic values of A are distinct. For the case in which there is a repeated
characteristic value and A is nonsymmetric, the determination of the
number of independent modal columns is not quite as clear. The reason
for the ambiguity is that there is no unique correspondence between the
order of a repeated root of the characteristic equation and the degeneracy
of the corresponding characteristic matrix [XiI — A].
If Xi is a repeated root of order p, the degeneracy of the characteristic
matrix cannot be greater than p, and the dimension of the associated
vector space spanned by the corresponding x2 is not greater than p. The
problem arises when there is a repeated root of order p, and the de¬
generacy q of [A*I — A] is less than p. Only q < p linearly independent
solutions to Eq. 4.7-12 can be found. The dimension of the associated
vector space for the xt is less than p, and p linearly independent char¬
acteristic vectors corresponding to A,- cannot be obtained. Only when the
(n x n) matrix A is symmetric is the degeneracy of [AJ — A] definitely
equal to p for a p-fold root, so that p linearly independent characteristic
vectors can be found.
For the case in which the degeneracy of [AZI — A] is equal to one
(simple degeneracy), the modal column can be chosen to be proportional
to any nonzero column of Adj [AtI — A]. This is the only column that
can be obtained for the set of p equal roots. For the case where the
degeneracy of [2*1 — A] is equal to q > 1, Adj [AtI — A] and all its
derivatives up to and including
{Adj m - A]}L=i(
Cl A
1 0“
0 1
-3 3
~ 1 -1 O'
[AI — A] |A=1 = 0 1 -1
-1 3 -2
is equal to one. Therefore there is only one modal column corresponding to the triple
root. Since
~1 -2 r
Adj [2,1 - A] |Ai=1 1 -2 1
1 -2 1
It can span only a one-dimensional space. The determination of two additional vectors
for this case is considered in the next section.
Repeated characteristic roots and multiple degeneracy.
>
A = 1 2 l -1 2-2 -1
II
1
_1
o
o
_0 0 i_
1
i
The characteristic values of this matrix are 1, 1, 3, a double root at 2 = 1 and a single
root at 2 = 3. The degeneracy of [2,1 — A] for 2t- = 1 is equal to two. Since the charac¬
teristic matrix has “full degeneracy,” it is possible to obtain a linearly independent
vector solution for each of the repeated roots.
The adjoint matrix is
-(2 - 2)(2 - 1) (2 - 1) (2 - 1)
Adj [21 — A] — (2-1) (2 - 2)(2 - 1) (2 - 1)
0 0 a-m - 3)
Substitution of 2, = 1 in any column of the adjoint matrix yields a null column. The
first derivative of the adjoint matrix is
"22-3 1 1 "
d
{Adj [21 - A]} 1 22-3 1
1).
0 0 22 -4
Sec. 4.7 Characteristic Values and Characteristic Vectors 239
"1 -1 0 “
1 -1 1 = 1 1
_0 0 1_
0 0 — 2_ 0 — 2_
Hence the two modal columns corresponding to the repeated root at A = 1 are given by
~-r " r
Xl = i Xo = l
o_ 2_
The characteristic vector or modal column for A,- = 3 can be chosen to be proportional
to any nonzero column of the matrix Adj [A,I — A], Thus
The case where the matrix A is real and symmetric occurs so frequently
when dealing with linear circuits that special attention should be paid to
the form of the characteristic values and characteristic vectors associated
with such a matrix.
A fundamental property of real symmetric matrices is that the char¬
acteristic values of a real symmetric matrix must be real. This can be
shown by assuming that the characteristic values are complex. They must
then occur in complex conjugate pairs, as must the characteristic vectors.
Thus Ax = Ax and A*x* = Ax*. Premultiplying the first equation by
x*r and postmultiplying the transposed form of the second equation
by x gives A(x, x) = (x, Ax) and A*(x, x)= (x, ATx>. Noting that,
for a symmetric matrix A = AT, these expressions require
(A — A*) (x, x) = 0.
A2X1TX2 = XxTAx2
Example 4.7-4. Show that the characteristic vectors of the symmetric matrix A are
orthogonal, where
"-3 2~
A =
2 0
and x2
t This is always the case if the characteristic values of A are distinct, or in the case of a
symmetric A matrix. See Reference 7.
Sec. 4.7 Characteristic Values and Characteristic Vectors 241
. . . • • •
S3
a in X11
rH
rH
al2 X12 Xln
• • • • • •
— *21 a22 a2 n X21 x22 X2n
(4.7-16)
• • • • • •
anl 0' n2 @nn Xnl Xn 2 Xnn_
or
MA = AM
where
A = M_1AM (4.7-17)
y = AMq (4.7-19)
In terms of the new coordinate system ql9 q2, q2, ... , qn, the set of equations
described by Eq. 4.7-21 is uncoupled. Note that the qi coordinates are
in the direction of the characteristic vectors. These coordinates are called
the normal coordinates of the system. By transforming to the normal
coordinates, the characteristic values, and hence the modes of the system,
are isolated. It is for this reason that M is called the modal matrix.
Recognition of the characteristic vectors as uncoupling the coordinates
forms the core of the mode interpretation of linear systems.
The columns of the modal matrix M form a basis, and the rows of
M_1 form a reciprocal basis, in the original space. If the columns of the
modal matrix are called ul5 u2, . . . , un, and the reciprocal basis is denoted
by rl9 r2, ..., rn, then any vector y can be expressed as
To illustrate the equivalence of this form and the use of the normal
coordinates, consider
Z2
Example 4.7-5. For each of the cases below, show that the product M_1AM is a
diagonal matrix with its elements equal to the characteristic values.
Distinct roots (see Example 4.7-2).
M
1 1 1 M = 1 1 l 0 2 -2
S!
II
1 3 -1 1 -14 i_ 15 3 12_
Consequently,
"1 0 0 "
M AM 0-2 0
0 0 3 /
“1 0 0 ~
M_1AM = 0 1 0
0 0 3
This procedure yields a diagonal form when the degeneracy of the characteristic matrix
[A J — A] is equal to the order of the root. If it is not, a similar procedure yields a more
general Jordan form discussed in the next section.
Example 4.7-6. For the linear transformation given, show that the columns of M form
a basis and the rows of M-1 form a reciprocal basis. Express y in terms of the charac¬
teristic vectors.
f 0 11
A2 + 3A + 2 = 0
244 Matrices and Linear Spaces
The roots of this equation are 2X = — 1 and 22 = —2. The adjoint matrix is
2 + 3 1
Adj [21 - A]
-2 2
The characteristic vectors are
" r ■ r
Xi = and x2 =
-1 -2
1/V2' 1/V5
Ui and u2 =
-1/V2 -2/V5
The modal matrix is
1/V2 l/v%'
M =
-l/v7! -2/V5
The inverse modal matrix is
2^2 V2
M1
— V5 -V5
2V2" -V5
_1
and r2 =
1 <N
L. _
<
>
Ul
1
1
such that
<rl5 Uj> = 1 <r2, uj> = 0
x2
If the characteristic vectors u* are used as a basis, then the vector y can be expressed as
_*31 *32 * _
33 _0 1 0_ 31 *33 *32
Operation 2. If k times the y'th row is added to the zth row of A, this is
equivalent to premultiplication of A by Q2, where Q2 is the (n x n) unit
matrix with the element k in the zth row and y'th column (z y). In a
similar fashion, the addition of k times the zth column to the yth column
is equivalent to postmultiplication of A by Q2. For example,
0 k 1_ _*31 *32 * _
33 _*3i + kan *32 ~F ^*22 *33 + ka23_
"*11 *12 * _
13 '1 0 O' "*n *12 + ^*13 *13
_*31 *32 * _
33 _0 k 1_ _*31 «32 + ktf33 * _
33
246 Matrices and Linear Spaces
_i
k_ ka33_
a
_0 0 _031 032 033_ ka%2
CO
rH
0n 012 013 "1 0 0"
~0n 012 ka13
0 1 0 — ko 23
021 022 023 021 022
Equivalent Matrices
B = PAQ (4.8-1)
Two matrices which are equivalent have the same rank. Conversely, it
may be shown that two (m x /?) matrices are equivalent if and only if they
have the same rank.
Normal Form
where Ir is the (r x r) unit matrix. These forms are called normal forms,
or canonical forms. Note that, if the nonsingular matrix A can be reduced
to a unit matrix by a sequence of operations on the rows of A, then
P = A-1, Q = I. This is another method of finding A-1. In general,
if A is reduced to a unit matrix by a sequence of elementary operations,
then
A = P !PAQQ 1 = P-XIQ 1 = P !Q 1 (4.8-2)
A =
1
1
r-H
O
~—2 —3"
_i
O
2 —3_
L
1
_i
r-H
O
o
_1
i
i
1__1
i
i
o
<N o
o
1 0“
1
1
_i
r-H
r-H
O
H-k
o
L
1
This reduction was performed using only row operations, in order to illustrate how
the inverse of a matrix may be obtained. The sequence of elementary operations was
O' 1 3' 0 r
0 1 0 1 1 0
The product PA is
in
i_
i
o
“_3
2 ~1 O'
_i
—i
o
1 2 —3_
i
Therefore P = A-1.
Singular matrix.
2 —3'
2 -1
-3 4
_0 -1 0 _0 0 l X) -1 0_
Step 5. Add (—2) times the first column to the second column.
i o on
_j
1_
T—H
o
o
o
o
I2 i
0 1 0 0 1 0 —
0 1 0 = “ |°
i_
-1
o
o
o
0 1 1_ 0 0_ _0 0 0_
r
1 1 0 -1 2 -1 0 1 1 = 0 1 0
ii
_5 1 4_ -1 -3 4_ 0 0 1_ _0 0 0_
Sec. 4.8 Transformations on Matrices 249
Collineatory (Similarity):
B = Q_1AQ, or P = Q 1
Orthogonal:
B = QtAQ = Q_1AQ, or P = QT = Q 1
Congruent:
B = QrAQ, or P = QT
Conjunctive:
B = Q*rAQ, or P = Q*T
Unitary:
B = Q*rAQ = Q'AQ, or P = Q*T = Q1
y = Ax (4.8-3)
x' = Px X = P-1x'
, , , (4.8-4)
y' = Py y = P V
The relationship between y' and x' in the new coordinate system is to
be found. To obtain this relationship, premultiply both sides of Eq.
4.8-3 by P, forming Py = PAx. Using Eq. 4.8-4,
The matrix B, relating x' and y' in the new coordinate system, is obtained
from A by a similarity transformation.
Similarity transformations have the extremely important property
that the characteristic values are invariant under such a transformation.
To show this, let B = PAQ. Then
Since P = Q_1, the product of the determinants |P| and |Q| is equal to
unity. It follows that
|B - AI| = |A - AI| (4.8-7)
Orthogonal Transformation
x = Qx', P 1 = Q
X TX = xTx = xTQTQx
This requires that QTQ = I, or
Qr = Q1 (4.8-8)
Therefore, if a transformation from one mutually orthogonal basis to
another mutually orthogonal basis is made, the transformation matrix
Q which relates a vector in the new coordinate system to a vector in the
original coordinate system must satisfy the relation shown in Eq. 4.8-8.
The transformation is then called an orthogonal transformation. The
matrix Q is called an orthogonal matrix. An orthogonal transformation
is a special case of a similarity transformation. Lengths and angles are
preserved.
Sec. 4.8 Transformations on Matrices 251
Unitary Transformation
Congruent Transformation
Two matrices A and B are called congruent if one is obtained from the
other by a sequence of pairs of elementary operations, each pair consisting
of an elementary row transformation followed by the same elementary
column transformation. As a consequence of the definitions of elementary
transformations, it follows that two matrices are congruent if there exists a
nonsingular matrix Q such that
B = QtAQ (4.8-11)
Congruency is a special case of equivalence, so that congruent matrices
have equal rank. A congruency transformation is a transformation to
a new basis such that, if two vectors x and y are related in the original
basis by Eq. 4.8-3, the vectors x' and y' in the new basis are related by
the equation
y' = Bx' = QTAQx (4.8-12)
Transformations of this type are useful when dealing with quadratic
forms, which are discussed in the next section.
252 Matrices and Linear Spaces
Conjunctive Transformation
Congruent Transformations
0 0 0
The integer p is called the index of the matrix, and the integer s—p —
(r — p) — 2p — r is called the signature of the matrix. Two (n x n) real
symmetric matrices are congruent if and only if they have the same rank
and the same signature or index.
Example 4.8-2. Reduce the given symmetric A to the canonical form of Eq. 4.8-15.
2 2 ~
3 5
5 5
Sec. 4.8 Transformations on Matrices 253
"1 0 0“ “1 2 2~ "1 2 2~
0 1 -1 2 3 5 —
0 -2 0
_0 0 1_ 2 5 5_ 2 5 5_
Step 3. Subtract twice the first row from the third row.
0 -2 0 0 1 0 = 0 -2 0
_0 0 1 _ _0 0 1_ _0 0 1
Step 5. Multiply the second row by 1/V 2 and the second column by 1/V2.
0 1/V2 0 0 -2 0 0 1/V2 0 = 0 -1 0
_0 0 1 _ 0 0 1 _0 0 1 _ _0 0 1_
Step 6. Interchange the second column with the third column and the second row
with the third row.
“1 0 0 ~ "1 0 0 " "1 0 0 “ '1 0 cr
0 0 1 0 -1 0 0 0 1 =
0 1 0
_0 1 0 _ _0 0 1_ _0 1 0 _0 0 -1 _
In this case the index of the matrix \sp = 2, the rank is r = 3, and the signature is 5 = 1.
(4.8-17)
_0 0 0_
The index p and signature 5 are the same as defined for a real symmetric
matrix.
Skew-Hermitian Matrices. An (n x n) skew-Hermitian matrix of rank
r can be reduced to the canonical matrix of Eq. 4.8-19 by a conjunctive
transformation.
./i, ! 0 0
B = Q*'AQ = 0 j^-r—p 0 (4.8-19)
_ 0 0 0_
Similarity Transformations
1
K 1
h
A1
i
h i
h i
K i
The result is that it is not clear whether the Jordan form given above or the
form
K l
Ax
Ax 1
J =
Ax
A2 1
A2
^ 2) "T ^222 — 0
qnqzitti - K) + ^2i2 = 0
If Ax = A2, then q22 and q2X must vanish. However, if these terms vanish, then |Q| = 0,
which violates the similarity transformation. Therefore the matrix A cannot be
diagonalized by a similarity transformation if Ax = X2. Since A is already a Jordan
form, no further transformations are considered.
Example 4.8-4. Reduce the A matrix of Example 4.7-3 with repeated characteristic
roots and simple degeneracy to Jordan form. Determine M.
Since
"0 1 0 “
A = 0 0 1
1 -3 3
has a characteristic value A = 1 of order three and only simple degeneracy, there is only
one linearly independent characteristic vector. Hence the Jordan form contains one
Jordan block. Also, the order of A, minus the degeneracy indicates two “ones” in the
Jordan form. Hence the Jordan form must be
'1 1 0"
0 1 1
_0 0 1
*21 — 3-11
*31 = *21
where xxx, x2X, and x3X denote the elements of xx. These equations yield *xx = x21 = a?3X.
Thus
T
xx = 1
1
is a characteristic vector. Note that it must be, and is, the same characteristic vector
determined in Example 4.7-3.
Now, considering x2, the second of Eqs. 4.8-22 gives
Substitution of the first two equations into the third yields &12 — 3'12* Hence x12 is
arbitrary. Let x12 = 1. Then
^33 = “b 2
Again, substitution of the first two equations into the third yields an arbitrary com¬
ponent, since it gives a;13 = x13. Hence let a;13 = —1. Then
~-r
x3 = o
2_
Thus
_i i -r
M = 1 2 o
_1 3 2_
Asa check,
’ 4 -5 2“
M-1 = -2 3 -1
1 -2 1
and
1 O'
M^AM 1 1
0 1
Example 4.8-5. Reduce
"1 0 O'
A = 1 1 0
2 3 2
to Jordan form and determine M.
Evaluation of |2I — A| yields the characteristic equation (4 — 1)2(2 — 2) = 0.
Considering the degeneracy of [XV — A] for 2 = 1,
0 0~
[I - A] = 0 0
-3 -1
is of degeneracy one. Hence J consists of two Jordan blocks. One is a first order block
corresponding to 2 = 2. The second is a second order block with a single “one.” It is
'2 0 O'
J = 0 1 1
0 0 1
Sec. 4.8 Transformations on Matrices 259
x3
Thus
0 0
M = 0 1
1 -3
As a check,
_5 3 r
M1 0 1 0
1 0 0
and
'2 0 0~
M AM = J = 0 1 1
0 0 1
Example 4.8-6. Reduce
"0 0 10 “
0 0 0 1
0 0 0 0
0 0 0 0
to Jordan form, and determine M.
260 Matrices and Linear Spaces
0 0 0 1 0 0 0 1
0 0 0 0 0 0 0 0
Equations 4.8-22 can be used on a trial and error basis to determine the correct form.
Assume that the correct Jordan form is the first gne given, i.e., one consisting of
(lxl) and (3 x 3) Jordan blocks. The first of Eqs. 4.8-22 gives
»31 = 0
xn = 0
0 = 0
0 = 0
Thus associated with the (1 x 1) Jordan block is
*^21
0
where xyl and x21 cannot both be zero, but are otherwise arbitrary.
Now, considering the (3 x 3) Jordan block, the first of Eqs. 4.8-22 gives
x32 = 0
X j2 =: 0
0 = 0
0 = 0
Hence
X12
X22
x, =
0
0
where x12 and x22 cannot both be zero and must be chosen so that xx and x2 are linearly
independent.
The second of Eqs. 4.8-22 gives
X22 = 0 +
x±2 — 0 +■ x.,2
0 = 0 + 0
0 = 0 + 0
Thus
x\z
X2Z
3? 2 2
Sec. 4.8 Transformations on Matrices 261
0 = 0+ x12
0 = 0+ x22
The last two expressions violate x2 0. Hence the proper Jordan form cannot be the
one consisting of (1 x 1) and (3 x 3) Jordan blocks.
Assuming now that the correct Jordan form consists of two (2 x 2) Jordan blocks,
the first of Eqs. 4.8-22 yields
*n
_ *21
Xi
0
_ 0 _
where xu and x21 cannot both be zero.
The second of Eqs. 4.8-22 gives
x32 = 0 + Xu
*42 — 0 + x21
0 = 0 + 0
0 = 0 + 0
Thus
*&12
*22
x2
xn
21
xi3 = 0
0 = 0
0 = 0
Thus
~xiz
x23
Xs= 0
0
where x13 and x23 must be chosen so that xx and-x3 are linearly independent.
The second of Eqs. 4.8-22 gives
x3i = 0 + x13
*41 0 + x23
0 = 0 + 0
0 = 0 + 0
262 Matrices and Linear Spaces
Thus
~xli xn X12 X13 xli~
_X23_ _ 0 X2 1 0 *^23_
where the components of xu x2, x3, and x4 must be chosen so that these vectors are
linearly independent.
As a simple check, let
*1 0 0 0 “
0 0 1 0 \
0 1 0 0
0 0 0 1
Then
ri 0 0 0“
0 0 1 0
M 1 =
0 1 0 0
_0 0 0 -1_
and
~0 1 0 O'
0 0 0 0
M XAM
0 0 0 1
0 0 0 0
There are many other canonical forms which can be obtained, the Jordan
form being a special case of the more general hypercompanion form.
A fairly complete listing of these forms is given in the literature.11,12
For most of the problems that the reader will face, knowledge of the
Jordan form is adequate.
where all components are real, is called a bilinear form in the variables
x., yj% This form can be written compactly as
n n
B = 12,2=1 3=1
(4.9-1)
Sec. 4.9 Bilinear and Quadratic Forms 263
or in matrix form as
_1
a 12
M
flu Vi
3
*21 a 22 a2n y2
B = [x1 x2 • = xrAy = (x, Ay)
Q = 12
i—1 3 = 1
auxixi (4-9-4)
Note that the coefficient for the term xyc^i ^ j) is equal to {aij -f au).
This coefficient would be unchanged if both aii and aH are set equal to
\{aij + ax). Therefore the matrix A can be said to be a symmetric matrix
without any loss in generality.
If the matrix A is a Hermitian matrix, such that a{j* = aH, then the
corresponding Hermitian form is defined as
The theorems which are developed for a real quadratic form have a
set of analogous theorems for the case of a Hermitian form. Since the
proofs of the analogous theorems require only minor changes from
the proofs for the real quadratic case, the latter theorems are stated
without proof.
Transformation of Variables
The linear transformation x= By, where B is an arbitrary (n x n)
nonsingular matrix, transforms the quadratic form of Eq. 4.9-3 into a
quadratic form in the variables yl9 y2, . . . , yn. This form is
Q = yTBTABy (4.9-6)
If the symmetric matrix A is of rank r < n, the modal matrix can still
be formed such that the transformation shown yields a diagonal matrix
with the diagonal terms equal to the characteristic values of A. For this
situation the modal matrix M is not unique. There are infinitely many
ways in which a set of m orthogonalized characteristic vectors corre¬
sponding to a characteristic value of order m can be chosen. Note that,
if there is a zero characteristic value of order m, then there are only n — m
nonzero terms in the quadratic form. The matrix A is then of rank
r = n — m.
Example 4.9-1. Reduce Q to a linear sum of squares, where Q = <x, Ax) and
1 1"
0 2
2 0
The characteristic values for A and associated normalized characteristic vectors are
given by:
1
1
-1
-1
2 ~
i
/2 = 4, U) = ~ 1
V6
1
0~
1
= -2, us = 1
1
Sec. 4.9 Bilinear and Quadratic Forms 265
The transformation
The reduction to a sum of squares can also be approached by the Lagrange technique
of repeated completion of the square. This technique is demonstrated as follows:
Vz =
Then
Q = y 12 — -3I/22 + 8 y32
1/V3 -i -2
B = 0 1 5
0 0 1
This is not an orthogonal matrix, and therefore the new coordinate system does not
have mutually orthogonal unit vectors. However, the congruent transformation
BrAB does reduce the quadratic form to a sum of squares.
The number of positive terms p is called the index of the quadratic form.
266 Matrices and Linear Spaces
rri — sj ; 04 ^2 i ( , 2, . . . , r)
1
W; = z,- / = (r + 1, . . . , n)
The latter form follows from the definition of the canonical matrix of
Eq. 4.8-18.
A = CTC (4.9-12)
1_
1
I-H
o
o
o
1
1
by a congruent transformation. In the latter case, C is of rank r, and A
is positive semidefinite.
The quadratic form is called positive semidefinite if it is non-negative.
It can be zero when the vector x is not zero. This case arises when A
is singular. From Eq. 4.7-8 it follows that, if A is singular and of rank r,
then it must possess n — r characteristic roots equal to zero. There are
then n — r terms of Eq. 4.9-8 which are identically zero, even when the
associated yt components are nonzero.
Analogous to the definitions above, the quadratic form may be negative
definite and negative semidefinite. The conditions required for these
forms, as well as the corresponding Hermitian cases, are listed in Table
4.9-1.
II
Ai — #n, A2 — = A|
CO
c = aad (4.9-14)
268 Matrices and Linear Spaces
s: S.
1/5
t- -o
<3
s= <
t.
/■-s
o
C < as 1
00 — > 1
C S ■4—* w ^ as '
■o —
as os
. ’c/5
< « A-S as 1)> as
O . > >
OS Co Cl Q.
<3 — •' ■*■*
- <3„ 00 «J • w
j o • '55 .as M
c C, oS • o . c « o c o
c d a . £ <1 0 <] D-<] D.
CU <3
<1 *— ~ o
,
iH
CO
- c
^ 03 4- C <3■ "E <\ a - °
^ V.
<3 <3 1 I I
St *4
St v> * *
*4 C4 st
* M
*4 V- d | d
M Si
5*
+
2
V
n
1 ;
+ : + ; : i1
: + : + d Cl
• d Cl M *4
p ^ *d *
d
l_ 1 1 *4 ci Cl N dd 14
O + 1 * Cl + * Cl
Ph Cl Cl C4 d Cl w
^ 1
5* + ^ + 1
14 *4
4- d 4 d fH *
1 *4 M
d * Cl * S*
^ M hT M^ 1 1
IIII III* IIII IIII 11 II
II II
Ol o» OJ o> O) O) Ol O)
as
-*—* a>
(U -4—*
d>
-4—» 'E -4—» s
«c c u
c as <D
6\ .. o U o •a r,° Uo
CJT
<D tT1 ° * o TD Po &H
* o
Tt <D
Q E W ii Q E U ll
< e~, A * A as II * D VA Va <D
0) U— U — c/3 u_ u_ > 1 — 1 — C/5
3 > ,1 u „ u as II u ,| u d>
a w II — II — > IIII —
u IIII —
u ”c3 II — II — >
c/5
H -4-*
A
A
O 1/5
Oh o z W)
as a. <u
4-* o
C/5 C/5 o
1- C/5
-*-»
*-»
<d
'£ 00 o o OS
<u Si On w O N >
w d o _ £ .> O —
A V t- rf a as
aS ^ •- <a S3 .t;
*■< V O JZ </5 3 .e oo
crt:
os o as
•2
E> cr *- o
.c | as o a. e
U
T3 K
aS ^ II V =: o s: o
as
^ -a Co, Co, II II V II
c c
as 1-1 II II ^ co.
P<
o
A
*[ o o o
II s:
A
1 <
o g as E o E o E
u.
2 2 1 £ 1 -2 E <2
■a c ■O C -n c e
aS aS OS 03 aS aS .5
E^E ■% a- E | or E ■■§
- S £ — o £ - E e £
aj ,0 t- c3 .O u- c3 .O U-. as c2 t-
(L) ^ a> cl> a> CL) ^ <D as a>
c* X Ot I 2
Sec. 4.9 Bilinear and Quadratic Forms 269
where A is a diagonal matrix whose diagonal elements are A1? A2, . . . , A„,
and D is a triangular matrix of the form
"1 d12 diz " * din
0 1 ^23 ^2 n
0 0 1 * * ‘ *4 n
_0 0 0 • • • 1
Substitution of Eq. 4.9-14 into Eq. 4.9-12 yields
A = DrAD
~ 1 0 0 ... Q 0 0 • • • 0“
pi
d\2 1 0 ... o 0 22 0 • • • 0
^13 ^23 1 ... o 0 0 2
a3 . . • 0
^ :
... ] _ 0 0 0 • • •
<4 n <4 n K.
3
1
~1 d\ 2 4.3 * • • din
0 1 ^23 • • * d2n
(4.9-15)
X 0 0 1 • * • <4«
0 0 0 ... 1
Q1 = Kv\ + + ' * * +
By a similar argument, the discriminant of this quadratic form An-1 is
equal to
An_i = -4^2 ' * ’ ^n-1 (4.9-19)
270 Matrices and Linear Spaces
K = Aj
a2
2*2
A3
a3 (4.9-21)
a2
Clearly, if all the elements are to be positive, then all the leading principal
minors A1? A2, . . . , An must be positive. Therefore a real quadratic form
Q = <x, Ax) is positive definite if and only if all the leading principal
minors of A are positive.
For the case of Hermitian forms, a set of analogous statements can be
made. A summary of the useful statements regarding real quadratic and
Hermitian forms is given in Table 4.9-1. The statements about the negative
definite forms can be proved by requiring (—A) to be the matrix of a
positive definite form.
Powers of Matrices
for scalar algebra. The matrix A0 is defined as the unit matrix of order n.
AkAm — Ak+Tn
(A^)m = Akm (4.10-1)
A0 = I n
(A"1)771 = A —m (4.10-2)
<2n aVi
_a21 #22.
Let
bn b12
B2 = = A
b 21 ^22
Then
bii2 T b\2b2j ^12(^11 T b22) all a12
This is a set of four nonlinear simultaneous equations which has no general solution.
A pair of numerical examples illustrate the ambiguity involved.
Let
"4 1
A =
0 1
Then b21 = 0, blx = ±2, b22 = dbl, and b12 — The square root of A is then
‘2 f
B = ±
0 1
As a second example, let
■4 O'
A =
0 4
One possible answer to this problem is
'2 O'
B = ±
0 2
272 Matrices and Linear Spaces
with 6X12 + 612621 = 4. Therefore there are an infinite number of square roots of A.
Matrix Polynomials
Note that the last term is multiplied by the nth order unit matrix In.
Example 4.10-2. Let N(x) — 3x2 + 2x + 1 and
Determine N(A).
N(A) = 3A2 + 2A + I
[4 41
r h
+ 2 2
ri °i rn
141
3
0 4 0 2 +
0 1
—
0 17
where A2, . . . , An are the roots of the polynomial N(x) = 0 and are
all assumed to be distinct. Similarly, the factored form of a matrix poly¬
nomial is
N(A) = pn(A - A1I)(A - X,l) • • • (A - Xn\) (4.10-5)
If the argument of the infinite series is replaced by the square nth order
matrix A, then the infinite series of A can be written as
Geometric series:
OO
A2 A3 Ak
exp A = I + A H-1-b ' ’ * H-b
2 3! k\
(4.10-8)
-A A2 A3 , (~l)fcA*
exp ( — A) = I — A H-b
2! 3! k\
Sine function:
A A3 A5 exP L/A] - exp [ — /A] ,41011)
sinA = A —-1-—
3! 5! 2j
Cosine function:
A2 A^ exp [j\] + exp [-jA]
cos A = I (4.10-12)
2! ^ 4! 2
274 Matrices and Linear Spaces
Trigonometric Identities
Cayley-Hamilton Theorem
then this generalization shows that the polynomial using A as the variable
is
N(A) = An + cqA^1 +-h cn_xA + cnl = MN(A)M-1
Wi)
N(X2)
M"1 (4.10-20)
WJ
where A1? X2, . . . , Xn are 4he zeros of the polynomial N(X).
If the polynomial chosen is the characteristic polynomial, i.e., if
N(X) = P(X), then = N(X2) = • • • = N(Xn) = 0. It follows that
A =
P(A) = A2 + 3A + 21
'-2 -3“ " 0 r "1 0“ "0 0“
+ 2 _
+ 3
6 7_ 2 — 3_ _0 1 _0 0_
Example 4.10-4. Find A-1 for the matrix of the preceding example by using the
Cayley-Hamilton theorem.
276 Matrices and Linear Spaces
A2 + 3A + 21 = [0]
or
A + 31 + 2A-1 = [0]
Therefore
A-1 = —iA - fl =
Reduction of Polynomials
v
r
A =
-3
A2 + 3A + 21 = [01 or A2 = -3A - 21
Consequently,
Similarly,
A3 = —3A2 - 2A = —3( —3A - 21) - 2A = 7A + 61
Hence
N(A) ( —15A - 141) + (7A + 61) + (-3A - 21) + A + I
-9 -10"
= — 10A - 91 =
20 21
Sylvester’s Theorem|
f The proof of this theorem closely follows the proof given in Reference 15.
Sec. 4.10 Matrix Polynomials and Infinite Series 277
n (a - /,!)
W = ’4-
IT (7 *,) 3=1
-
+.
4- c„[(A - AjIXA - A21) • • • (A - A^I)]
or
N(A)=ic,n(A-;,D (4.10-24)
fc=l 3=1
3 ^k
Since there is one factor missing from each of the product terms, N(A) is
clearly a polynomial of degree n — 1, with n arbitrary constants of com¬
bination. If the characteristic vectors of A are denoted by u1? u2, . . . , un,
then postmultiplying Eq. 4.10-24 by yields the relation
n
However, since Au* = Xl\xi or (A — AfI)uf = 0, all the terms except the
zth are zero. The zth term is not zero, since it does not contain the factor
(A — 40. Therefore
1_
I
«K
'*
•?>4
J1)_1
—) J- JL
L - 3 ¥=i
278 Matrices and Linear Spaces
fl 0, - A,)
2=1
3
Consequently,
ff (A - A,I)
2=1
n
N(A) = 2 AW ^-
n (Ai - a,)
2=1
0^3
This concludes the proof of Sylvester’s theorem.
Example 4.10-6. Calculate eA, using Sylvester’s theorem, for the A matrix of Example
4.10-5.
Since eA can be expressed as a convergent series in A, Sylvester’s theorem can be used
directly on eA, rather than on the infinite series representation of A. Certainly, if the
infinite series for eA converges, then eA can be determined by Sylvester's theorem.
Therefore
2
2 ^ z„a,.)
2=1
where
A - LI A - XJ.
= and Z0(A2)
— X-2 /, — Xx
Since
0
A =
-2
-1
2
Consequently,
le-1 - e-2 e-1 - e~2
Example 4.10-7. Calculate Ak, using Sylvester’s theorem, for the A matrix of the
previous example.
a* = 2
2=1
The sum of the contributions of all the roots with different values is then
N(A). Hence
1 { d3-1 fA(2) Adj (21 - A)'
N(A) = 2 (4.10-30)
T(s - 1 )\\dkS— 1 O. n -
3=1
3 A=A*
where the summation is taken over all the roots, with repeated roots taken
only once.
Equation 4.10-30 is the confluent form of Sylvester’s theorem. A
typical term of the summation, corresponding to a multiple root 2Z, can
be expanded into the form
s—1
1 d N(X) Adj (AI - A)'
(5 - 1)! i dfs—1 n (* - w
3=1
3 A=A,
zs.k(K) (4.10-31)
=2
k=i (/c — 1)!
where
dkN(X)
nw =
dXk X=X,
t A proof of the theorem for this form can be found in Reference 17.
280 Matrices and Linear Spaces
and
Adj (AI - A)
zM = : :
no- W
0=1
'o k=Xi
Example 4.10-8. Find the general form for any matrix function of A, where the
matrix function can be expressed as a matrix polynomial in
1 3"
0 2
2 4
Evaluation of
shows that the characteristic equation has a double root at A = 1, and a single root at
A = 2. The contribution to N(A) from the single root is
r -8 4 8‘
N(2) -22 11 22
2 -1 -2
Since
~2A — 4 1 3'
- {Adj [AI - A]} = 6 22-4 2
-5 2 22
"7-3 -5" "2-1 -3“ "9 4-8"
Zi(l) = 28 -12 -20 + -6 2 -2 _
22 -10 -22
_—7 3 5__ 5 -2 —2_ _—2 1 3_
“7-3 —5“|
Zo(l) = 28 12 -20
-7 3 5
" 7 -3 -5“
+ dN(X)
28 -12 -20
dA A=i
—1 3 5_
dN{A)
Nil) = e2h A/(l) = and = tel
dA ;.=i
and
N(A) =
' (9ef + 7tet - 8e20 (-4e* - 3tel + 4e20 (-8e* - 5tet + 8e2!) ‘
(22e; -j- 2Stet — 22e2i) (-10e* - 12tel + 1 le24) (-22c4 - 20tel + 22e2<)
_ ( — 2eJ - 7/e4 + 2e24) (e4 + 3/e4 - e24) (3e4 + 5/e4 - 2e24)
Cayley-Hamilton Technique
A =
The preceding technique is valid only for the case in which N(A) is a
polynomial function of A. When F(A) is desired, where T(A) is an analytic
function of A, in a region about the origin, an extension of the previous
method can be used. If F(X) is an analytic function in a region, it can be
expressed by an infinite power series in A, which converges in the region
of analyticity. Therefore the function F(A) can be expressed as a poly¬
nomial in A of degree n — 1. Consequently, the remainder R(X) of
Eq. 4.10-33 must be a polynomial of degree n — 1. It follows that, if
2(A) is an analytic function of A in that region,
T(A) = g(A)P(A) + R{X) (4.10-35)
where P(X) is the characteristic polynomial of A, and R(X) is a polynomial
of the form
R{X) = a0 + ajA -f a2A2 + • • • + a^A*1-1 (4.10-36)
The coefficients a0, als . . . , a„_]_ can be obtained by successively sub¬
stituting Ax, A2, . . . , An into Eq. 4.10-35. Since P(kt) — 0, the equations
nh) = R(h)
T(A2) = R(A2)
(4.10-37)
F(K) = R(L)
Sec. 4.10 Matrix Polynomials and Infinite Series 283
(4.10-38)
fi(A) =
PQ)
is an analytic function of X. Since the zeros of the denominator of Q(X)
are also zeros of its numerator, the function Q(X) is analytic in the region
of analyticity of F(X). Therefore Eq. 4.10-35 is valid for all values of X
in the region of analyticity of F(X). Consequently A may be substituted
for the variable X, if the region of analyticity includes all the characteristic
values of A. This substitution yields
F(A) = Q(A)P(A) + R(A) (4.10-39)
Since A is a second order matrix, the polynomial R(2) is of first order, i.e.,
R(X) = a0 + axA
Therefore the two linear equations obtained by substituting Xx and X2 into Eq. 4.10-35
are
F(Xx) = R(Xx) F(X2) = R(X2)
€ai< = a0 + a:Ax eA2f = a0 + ccxX2
ax = e 1 — e 2t
Hence
'a0 0 ‘ 0 ax"
F(A) = = a0I + axA = +
0 a0 —2ax — 3ax
2e-« _ e-2f €-< _ e-2«
Example 4.10-11. Determine eAf, where A is the matrix used in Example 4.10-8.
This matrix has a double root at 2 = 1 and a single root at 2 = 2. Since this is a third
order matrix, the polynomial R(X) is R(X) = a0 + ax2 + a222. The three equations for
the a*’s are given by F(2X) = and
dF(X) dR(X)
dX A—A^ dX A= +
where 2X = 1, and by F(22) = F(22), where 22 = 2. Thus the a’s are specified by
1
1_
*1 i r a0
w
td = 0 1 2 ai
_1
-1
d1
8
<M
i
_a2_ _1 2 4_ _e2t_
It is instructive to solve this set of three simultaneous equations by using the Cayley-
Hamilton theorem to find the inverse of the coefficient matrix. The characteristic
polynomial of the coefficient matrix is 23 — 622 + 42 — 1 = 0. Hence
T 1 1
5C2 + 4C - I = [0] where C = 0 1 2
J 2 4
Then
" 0 -2 1"
-i = C2 - 6C + 41 = 2 3 -2
-1 -1 1_
Therefore
a0 = -2/6* + € it
-it
«i = 2d + 3 td - 2i
a2 = — d — td + €2*
Hence
eA t — a„I + ajA + a0A2
where
"013" ' -9 6 14"
A = 6 0 2 and A _ -10 10 26
_ —5 2 4_ _ -8 3 5_
Sec. 4.10 Matrix Polynomials and Infinite Series 285
The result is
" (9e* + 7/C - 8e2f) (—4ef - 3Ul + 4e2() (-8e‘ - 5/C + 8e2t) "
eA^ = (22ee + 28/C — 22e2') (-IOC - 12/e' + 11C) (-22e' - 20/e' + 22e2')
_ (-2et - 7tet + 2e2') (ef + 3/C - e2') (3ef + 5/e' - 2e2')
This matrix checks with the result obtained in Example 4.10-8 using Sylvester’s
theorem. A considerable amount of labor is involved in either method, but this is
usually the case when dealing with a matrix of order higher than two. Generally, the
Cayley-Hamilton technique requires much less labor than the use of Sylvester’s
theorem.
... JK— 1~
'FOX -1 Ax a0
}n — 2
F(l2) 1 L a\ . . . Ag ai
Jn—1
_F(K). J K K • • • An J J^-n— ]_
was analyzed, and the modal matrices M and M-1 were found to be
0 -2
From the generalization of Eq. 4.10-20, F(A) = MF(A)M x. If F(A) = eA/ is desired,
then
r* 0
F(A) = M M
„-21
0
286 Matrices and Linear Spaces
~ 2e_< - e~2t
e\t —
_ — 2(e_< - e-2t) — (e-i - 2e-20.
A2/2 A3/3
eA* = exp (At) = 1 + A t + — + — +•• • (4.11-1)
2! 3!
This series is absolutely and uniformly convergent for all values of the
scalar variable t. The derivative of the exponential function eAt with
respect to t is then the term by term differentiation of Eq. 4.11-1, or
Often the situation arises where the polynomial operator N(p) must
operate on the matrix product eA*B(t). It is assumed that the product
AB(t) exists, but that the product B(7)A does not. In this case,
In general,
/[€A*B(0] = eAt(pI + A)fcB(f) (4.11-5)
Consequently,
A(/?)[eA<B(0] = eAtN(pI + A)B(t) (4.11-6)
t ft ft a v c* a¥
ea' dt I dt + \ Atdt + \ ——dt + — dt +
o Jo Jo 2! Jo 3!
_ , At2 , A2t3 , AV ,
It +-h-b-+*••
2! 3! 4!
Hence
At
A I €A* dt = eAt - I
-1/ A t
eAi dt = A-\eAt At
- I) = (eAi - I)A-1 (4.11-7)
o r
Example 4.11-1. Find eM dt, where A =
'o ■2 -3
SX
II
n\
Jo
-
Tyv
<N
_ 2e-‘ e~2t - 1
1 yv
l
l
288 Matrices and Linear Spaces
-f -¥
A-1 =
1 o
le-1 - e- 21
~
_1
_1
toH
3
to
1
1
n\
<T\
-¥
1
1
■I 2
A~xeAt — A'1
1 0 -2c~l + 2e — it e~* + 2e~2t 1 0
3e.-it .-21
■2e~* + — + -
2 2 + v
2 + \2
2e-‘ - e -21 - 1 €-* - €-2t
exp A(A) dX
L Jo
oo i r (*t
exp A(A) dX = 2 - A(X) dX (4.11-8)
k=o k! Uo
Example 4.11-2. Find
7 O'
exp A(A) dX if A =
L Jo 0 t
- 0
2
A(A) dX =
r2
° 2
Consequently,
~t2 k ~€‘2/2 0 '
0
00 1 2
exp A(X) dX
T\ t2
L Jo
0
2 0 et2/2
_a_
dx9
(4.11-10)
Jdx,
L_^„_
d_
dx1
3
dXr
gradx)(grady =
-dVi dy2 dym-i
dx n_
2
Note that
1__i
<N ■n-
'1 3" xx + 6x2
iH
2 Ax = 2
_1
_x2_ + 6xx_
(M
i
If both A and the variables xl9 x2, . . . , xn are functions of time, then
the derivative of the quadratic form with respect to t is given by
21
A(0 =
It 1
Since Q(t) = t2xx\t) + x2\t) + 4to1(r)a;2(r), the derivative of Q(t) with respect to t is
~t2 2f
2<x(7), A(f)x(/)> = 2[xx{t) x2(t)]
21 1 js2(0_
= xx{t)[2t2xx{t) + 4tx2(t)] + x2(t)[2x2(t) + 4txx{t)]
and
~2t 2“ *i(0
<x(t), A(t)x(0> = MO *2(01
1_
_1
N>
«M
i
= 2 txx2(t) + 4x1(t)x2(r)
The scalar product of two real-valued functions f(t) and g{t) over the
interval {a, b) is defined as|
</, g) = f dt (4.12-1)
Ja
Norm of a Function
Orthogonal Functions
Two functions f(t) and g(t) are orthogonal over the interval {a, b) if
their scalar product vanishes, i.e.,
</.*> = 0 (4.12-4)
A set of normalized functions (f>i(t), • • • is said to be an orthonormal
set if the members of the set obey the relation
<&.&> = «« (4-12-5)
Similar to the approach used in the Gram-Schmidt orthogonalization
procedure, a set of n linearly independent functions can be used to derive
a suitable orthonormal set of functions.!
/(dt = 2 ak4>k(0
A=1
Since </, </>3) = and (</>_,-, <£fc) = <5^, it follows that c3- = sq. Therefore
the coefficients ak should be adjusted to the expansion coefficients ck.
This approximation by means of a minimization of the mean square
error is known as an approximation “in the mean.”
Since <0 cannot be negative, it follows that
n
f\t) dt ~^ck > 0 (4.12-8)
&= 1
Equation 4.12-8 is known as Bessel’s inequality. Since the terms in the
approximating series are orthonormal, the addition of orthonormal
terms cj>n+1(t), </>n+2(0> • • • must decrease the mean square error between
the function and the approximation. However, even though the sum-
00
00
f(t)= lim 2
294 Matrices and Linear Spaces
Certainly the series converges to fit) in the mean, such that the mean
square error over the interval {a, b) tends to zero. The series does represent
f(t) at a given point if f(t) is a continuous function throughout the
interval, and if the series converges uniformly in the interval (a, b). How¬
ever, even when a complete set of orthonormal functions is available, the
convergence of the series is a rather involved problem and is not treated
here.18
These polynomial functions are orthonormal with respect to the weighting function
since
ck =
References 295
If, for example, the function /(t) = e t is to be represented by a Laguerre series, the
expansion coefficients of the first five terms are
CO Cj J, C-2 g, C3 — i6, C\ — 32
Thus
4
fit) = 2 CkLk = H — T it + y2 — 3V3 + TFsff4
*=0
The square of the norm of the function relative to the weighting factor is
poo poo
S = ml -14
k-0
As each orthonormal term is added to the approximation, the mean square error is
decreased, as shown in Table 4.12-1.
Table 4.12-1
A calculation of/(0 for a given value of t illustrates an interesting point. The approxi¬
mation for f(t) as each orthonormal term is added is also given in Table 4.12-1 for t = 2.
The actual value of f(t) at t = 2 is e~2 = 0.1353. Note that the error at t = 2 actually
increases as the fourth term is added, and then decreases as the fifth term is added.
This is in contrast to the mean square error over the entire interval, which decreases as
each term is added. A Maclaurin series expansion for f(t) at t = 2 is extremely poor,
as shown.
REFERENCES
Problems
r
1
Cx|
CO
"l
A = B 0
3 2 1 i
4.2 What are the conditions on the elements atj and bij of the (2 x 2) matrices
A and B such that AB = BA?
4.3 Compute AB, where
1 2 1 2 1 2 1 2
0 1 2 1 2 1 0 1
B =
0 0 0 1 0 1 0 0
0 0 1 0 1 0 0 0
Problems 297
4.7 Find the value of the following determinant by using only the definition
of a determinant.
-7 -4 3 2
3 2 -5 2
IA
6 4 0 -4
6 4 1 -5
4.8 Find all the minors and cofactors of the determinant given in Problem
4.7. Show that the Laplace expansion of the determinant along any row
is equal to the expansion along any column.
4.9 Find the value of the determinant shown in Problem 4.7 by the method
of pivotal condensation.
4.10 Show that the product of the determinants |A|, |B|, is equal to the de¬
terminant of the product, |AB|, where
3 0 2 "l -1 4“
A = -2 -1 -1 B = 2 3 0
-1 -3 5 5 0 2
x2 — 1 X — 1 1
|A| X4 X3 2x -f- 5
X + 1 X2 X
n
A
'I
#in 0 0 • • • 0
hi ^12 a13
hi @n2 hn 0 0 0 • • • -1
1 0 0 • • • 0 0 0 0 • • • 0
0 1 0 • • • 0 0 0 0 • • • 0
0 0 0 • • • 1 0 0 0 • • • 0
0 0 0 • • • 0 1
Using the principles of pivotal reduction, the array to the left of the
broken line is eliminated. The inverse matrix appears in the lower
right-hand box, where each element is divided by the element in the
2n + 1 row and n + 1 column. Find the inverse matrix of
2
A =
1
by pivotal reduction and verify the result using A 1 = Adj A/|A|. Why
does this technique work?
4.19 Repeat Problem 4.16 using the technique of pivotal reduction.
4.20 Find A-1 and (d/dt)[A-1] for
2t - 2 t + 2 -3
A = 31 - 1 t -1
-1 At — 3 -t + 1
1 2 3 “2 3 4~
A = 2 4 5 B = 4 3 1
1_
1 2 4
i
4.24 Show that, if the symmetric matrix A is nonsingular, then the inverse
matrix A-1 is also symmetric.
Problems 299
4.25 Find the inner and outer products of the following pairs of vectors:
1 -1
1 and y = 1
2 1
1 +j -1
x ■i -j and y = 1 +j
2 +J2 1 ~j
2
-1
and y =
1
1
2+y3 1 -f- j 2
1 -/2 0
x and y =
3 1 +y
0 2
4.26 Find the value of a which makes ||x — ay|| a minimum. Show that
for this value of a the vector x — ay is orthogonal to y and that
IIx — ay ||2 + || ay ||2 = ||x||2. The vector ay is called the projection of
x on y. Draw a diagram for the case where x and y are two-dimensional.
4.27 Find the projection of the vector (1, 1, 1) on the plane
xx + 2x2 + 3x3 — 0.
4.28 Assuming a three-dimensional space, prove that the four vectors (1,0, 0),
(0, 1, 0), (0, 0, 1), and (1, 1, 1) form a linearly dependent set, but that any
three of these vectors form a linearly independent set.
4.29 Express the vector y = (6, 3) in terms of the basis vectors xx = ( —1, 2),
x2 = (1, 3). Determine the reciprocal basis.
4.30 Using as a set of basis vectors xx = (1, 1, 1), x2 = (1, 0, 0), and x3 =
(0, 1, 0), find an orthonormal basis by use of the Gram-Schmidt orthog-
onalization procedure.
4.31 What is the dimension of the vector space spanned by the following sets
of vectors ?
(a) xx =(1,2,2, 1), x2 =(1,0,0, 1), x3 = [3, 4, 4, 3]
(b) Xj =(1,1,1), x2= (1,0,1), x3= (1,2,1)
(c) Xj =(1,0, 1,0), x2 = (0, 0, 5, 0), x3 = (10,0, 1,0),
x4 = (5, 0, 7, 0)
(rf) xx = (1,0,0), x2= (0,1,0), x3 = (0,0,1), x4= (1,1,1)
300 Matrices and Linear Spaces
_1
_i
-'t
VO
IT)
1
1
i
i
4.33 Find the degeneracy of the product AB where A and B are the matrices
given in Problem 4.32.
4.34 Let Sx be a subspace of the ^-dimensional space S. Show that any vector
x which is not in S± can be represented by
x = y + yi
where y is a vector in and yx = x — y is orthogonal to all vectors in
Sv Draw a three-dimensional picture illustrating this theorem. (See
Problem 4.26.)
4.35 Show that, if the ^-dimensional vector x is orthogonal to a set of basis
vectors of an /i-dimensional space, then x = 0.
4.36 Solve the following set of equations using Cramer’s Rule.
+ x2 + x3 + = 10
xx
3x1 x2 — x3 — 4x4
+ —14 =
2x1 + 2x2 — x3 + x4 = 7
x± + 3x2 + 4x3 — xi = 15
4.37 Find the solution to the equation y = Ax using the inverse matrix A-1,
where
1 2 3" x1 ”f
A = -1 0 2 x = x2 y = 3
0 1 0 x3 5
0 0 • • •
1
0 1 0 • • •
0
0
0)
o|
1 n + 1
0 0 0 • • • 0 1J
Problems 301
Choose any element amk as the pivot term, where m and k are less than
or equal to n. Using the method of pivotal condensation, the rows and
columns of this array are reduced one at a time until only column n + 1
remains. This column consists of n + 1 elements clt c2, . . . , cn+1. The
solution for the k\h unknown xk is then given by
Cfc
xk = -
Cn+1
2x1 — x2 + x3 = — 1
(a) x2 + x2 -f x3 + x4 = 0
x± + x2 + x3 — x4 = 4
X1 + X2 ~ X3 + ^4 = —4
xx — x2 + x3 + £4 = 2
(c) 2x1 + x3 — x4 = 0
xx + 3x2 + 2x3 + 4x4 = 0
xx + x2 + x3 + x4 = 0
4.40 Find the characteristic values and characteristic vectors for the following
matrices.
3
1 2
1 3
(a) 0 2
1
-1 1
2
4.41 Find the characteristic vectors, the modal matrix, and the diagonal form
for the matrices
0 1 o " ~2 -2 3 ” 7 4 -f
(a) 0 0 1 (b) 1 1 1 (c) 4 7 -1
-2 -4 -3 1 3 -1 -4 -4 4
302 Matrices and Linear Spaces
have the same characteristic values but are not similar matrices.
4.43 If a characteristic equation /?(2S) = 0 with a multiple root of order s
has degeneracy*/, whereq > 1, prove that the adjoint matrix Adj [ASI — A]
and all its derivatives up to and including (da~2/dkq~2)[Adj (21 — A)]a=As
are null.
4.44 The matrix
0 0
1 0
4 1
0 0“
0 e2 0
0 o e3
1 0 1
0 0 2
A =
0 5 1
5 1 0
4.46 If the «th order matrix A has a characteristic value At- repeated q times,
show that the rank of — A is not less than n — q, and that the as¬
sociated invariant vector space (null space) is not greater than q.
4.47 Prove that the characteristic values of a unitary matrix [A-1 = (Ar)*j
and an orthogonal matrix (A-1 = AT) have an absolute value equal to
unity.
4.48 Prove that, if Xi is a nonzero characteristic value of A, then i A|/Af is a
characteristic value of Adj A.
4.49 If the matrix A has characteristic values A1? . . ., An, show that
{a) Am has characteristic values Axm, . . . , Xnm.
(,b) A-1 has characteristic values 1/AX,. . . , \\Xn
(c) AT has characteristic values A1? . . . Xn.
0d) kA has characteristic values k\, . . . , kln.
Problems 303
4.50 If A is a triangular matrix, such that the elements either above or below
the principal diagonal are zero, show that the elements of the principal
diagonal are the characteristic values of A.
4.51 A vector xk, for which (A — ^ 0, but for which (A — 2il)kxk = 0,
is called a “generalized characteristic vector” or a characteristic vector
of rank k associated with the characteristic value 2Z-. These generalized
characteristic vectors are useful in determining the modal matrix for
systems which have repeated characteristic values.
{a) Show that characteristic vectors of different rank are linearly inde¬
pendent.
(b) Find the generalized characteristic vectors for
1 0 0
A = 1 1 0
2 3 2
L-i +j i j
4.57 Introduction to Linear Codes: Suppose that four binary digits {c^, c2,
c3, c*}, where each c may be either zero or one, are to be stored in
a computer memory. However, one or more of the digits may be in¬
advertently changed during the read or write cycle. Thus, the numbers
read out of memory, {xx, x2, x3, x4}, may be in error. It is possible to al¬
leviate this situation somewhat by introducing “parity check bits”
c5, c6, c7 as specified below. The encoded data (cl5 . . . , c7) are then
stored in seven bits of memory.
Noise = €
Fig. P4.57
The system shown in Fig. P4.57 illustrates this procedure. c5, c6, and
c7 are computed from cl5 c2, c3, c4 in the encoder. The numbers xlt. .., x7
are later read out of the memory and passed through a decoder.
Cl
^2 e2 x3
Vi
c = € = X = y = V2
• . . J/3_
_C1_ _e7_ x7
x = c + e
0+0=0 00=0
0 + 1=1 0-1=0
1+0 = 1 1-0=0
1+1=0 1-1=1
0a) Show that the mapping from the x vectors to the y vectors is a linear
transformation T. Find the matrix A of this transformation T with respect
to the natural basis in V7 and the natural basis in V3 (the seven and three-
dimensional spaces, respectively).
(b) List the eight error vectors e which correspond to
(1) no error;
(2) an error in the ith place only for / = 1, 2, . . . , 7.
Note: x = c + e defines e
(c) Show that the vectors c are the null space of A. Find the matrix
of the form whose column vectors are a basis for this null space.
(d) If the only possible errors are those listed in (b), show that it is always
possible to detect which c has occurred by computing y. For each y list
the corresponding e.
(e) In general, any matrix of ones and zeros defines a “linear code.”
There are other 3 by 7 matrices which define linear codes which have the
same error-correcting properties as A. How many are there, and what
is their relationship to A?
(/) In general, if a code defined by an (m x n) matrix A of a linear trans¬
formation T from Vn into Vm is to correct for a possible set of n-
dimensional error vectors {0, el5 . . . , ep}, what condition must be placed
upon the transformed vectors . . . , T(ev)} in Vm? Prove your
answer using the linearity property for T.
(g) Is it possible to construct a single error-correcting code such as the
one above from a (3 x 8) matrix? a (3 x 6) matrix? Tell why, referring to
your answer to (/).
4.58 If the coefficient matrix A of the linear transformation y = Ax is singular,
then the null space of A is the vector space formed by the vectors which
satisfy Ax = 0. Find the null space of
1 2 -2
A = 1 3 3
0 1 5
306 Matrices and Linear Spaces
k\ (k2 &i)
0 k2
(d) Consider the transformation which is the result of first performing (b)
for k1 = V2 and k2 = 1/V2; and then performing («) for 0 =45°.
(1) What is the matrix of this transformation?
T
=£>
(2) Find the (second order) characteristic vector and the characteristic
value, and interpret in terms of (c).
4.60 A second order orthogonal matrix must be of the form
Make sketches of both (a) and (b), indicating clearly the directions of
all characteristic vectors.
308 Matrices and Linear Spaces
4.62 Find the Jordan form and a modal matrix for the following matrices.
4 2 2 0
-1 0 -1 -1
(«) (b)
-1 0 1 1
1 2 1 3
2 0 0 0 1 0 0 0
0 1 1 0 0 0 1 0
(c) id)
0 1 3 2 0 1 0 0
V
0 1 -1 2 0 -1 -1 -1
0 0 0 0
0 0 0 0
0 0 0 1
10 0 0
4.63 Determine if the following matrices are: (1) positive definite, (2) positive
semidefinite, (3) negative definite, (4) negative semidefinite
'2 1 l" 3 1 1
2 6 —3
1 3 0 1 7 1
0) 0b) 6 6 —3
1 0 1 1 1 -1
3 3
1 1 1
(e) 1 3 1
1 1 1
4.64 Consider the quadratic equation <x, Ax) = 1, x real, A(2 x 2), real,
symmetric with an > 0. The equation defines a curve in the x-plane.
(a) Show that, if A is positive definite, the curve is an ellipse.
(1) What significance have the characteristic vectors in terms of the
orientation of the ellipse?
(2) What significance have the characteristic values ?
(3) Draw a sketch to illustrate (1) and (2).
(4) What happens if there is only one characteristic value?
(b) Show that, if |A| < 0, the curve is a hyperbola. Repeat (1), (2), and
(3) above for the hyperbola.
Problems 309
A =
3. 5.
2 2
4.65 Consider the function Q(x) = (x, Ax>, where A is a real, symmetric
matrix.
(a) Show that x = 0 is a stationary point.
(b) Under what condition on A is x = 0 a relative maximum? Prove it.
(c) Under what condition on A is x = 0 a relative minimum? Prove it.
(<d) If A is nonsingular and fits neither of these conditions, then what is
x = 0?
(e) What happens if A is singular with either single or multiple degen¬
eracy ?
(/) Illustrate your answers to (b) through (e) by means of sketches of the
level curves of (1) to (4) below.
-2 r
(1) A = (2) A =
i -l
1 1 1
(3) A = (4) A =
1 1 1
4.66 For the matrix
r
-3
~0 1 o" “2 1 o"
(c) A = 0 0 1 (d) A = 1 2 0
1 -3 3 1 1 1
310 Matrices and Linear Spaces
~1 0 o' ~1 i o'
3
{a) A = 0 2 2 (b) A = 3 -1 0
0 2 2 4 4 1
3 3_
Prove that this equality exists if and only if g(x) = kf(x), where k is a
constant.
4.74 Prove that the Laguerre polynomials are orthonormal over the interval
(0, co).
4.75 One technique which has been proposed for linear process identification
using only the input-output operating record of the process is to model
the process with a set of orthogonal transfer functions. In this technique,
the output of the model of the process is taken as a linear sum of the
outputs of the orthogonal transfer functions. The input to the model is
the same as the input to the process. See Fig, P4.75. The coefficients
Fig. P4.75
Problems 311
ax, a2, . . . , ay are adjusted until the mean square difference between the
output of the process, y(t), and the output of the model, z(t), is a
minimum. The set of orthogonal transfer functions is then said to “model”
the process.
(a) If
N
z(t) = 2
1
show that the optimum settings for the coefficients are given by
lim — I y(t)Zj(t) dt
00 T time average of yiOz^t)
a, =
mean square value of Zj(t)
\t) dt
1
FIs) = i — 1
5 + 5V
®100 = -r-1-
+ Si
n—1
_no- sk)
$n0) = ^- , n > 2
IT 0 + Sk)
k=1
(c) If a weighting factor w(t) = e 2<xt is used, show that the orthonormal
transfer functions are given by
V 2(5-! + a)
#i0) =
s + s1
V 2(s2 + a)(5 — 2 a — ^
<f>20) =
(s + 51)(5- + S2)
n
_ ITO - 2a - sk)
®«0) = V2(sn + a) -- > n > 2
ITO + sk)
k—1
s — a + \s±\
3^0) = V2(a + aj)
(s + ax)2 + p*
s — a — Is-jJ
OgCi') = V 2(a + ax)
0 + a^2 + ft2
_ (J - a + W) n K* - 2a - a,)2 + p2]
022_i = V2(a + a^)- k~X -
I! K* + «*)2 + Pi?}
*=i \
0 - a - 1^1) XX [0 - 2a - afc)2 + Pk2]
/c=1
®2i = ^2(a + oti)
IT k* + **)2 + /Vi
fc=l
where i = 1, 2, . . ., n\2 and n = number of complex poles.
5.1 INTRODUCTION
Example 5.2-1. Draw a block diagram for the system governed by the differential
equation , . , ,
1 y 4- ay + by = v
y = v — ay — by
Integrating y twice, both y and y are obtained, as shown in Fig. 5.2-2a. The loop is then
closed by satisfying the requirement of the differential equation
y = v — ay — by
y y y
/ ->
/
(a)
(b)
Fig. 5.2-2
The only change that must be made to the previous block diagram is the addition of a
block which provides the v term, as shown in Fig. 5.2-3a. However, a block containing
differentiation is not generally utilized. Differentiators are noise-accentuating devices
and therefore are not employed in an analog computer simulation.
(a)
316 State Variables and Linear Continuous Systems
y — ev — dv y — ev — dv
/
(a)
dv
•• •• ,•
y — ev — dv
->
J ^6
(b)
\y~e\
) > jr
y “ ev
(d)
Fig. 5.2-4
Sec. 5.2 Simulation Diagrams 317
This can be simulated as in Fig. 5.2-3b. The output of the first integrator is y — v.
By adding v, y is obtained. The diagram is then completed as in the previous example.
In essence, the v input of Fig. 5.3-2a has been shifted to the right of the first integrator,
and the differentiation and integration operators cancel.
Example 5.2-3. Draw a block diagram for the system governed by the differential
equation
y + ay + by = ev -F dv + cv
y — ev — dv = cv — ay — by (5.2-1)
An additional term aev is returned to the input of the first integrator. The system can
still be simulated in the form shown, however, but with a modification of the blocks
containing the multiplying factors c, d, and e.
Assume, then, the form shown in Fig. 5.2-5a, where b0, blf and b2 are the proper
multiplying factors. Call the input to the first integrator q. The values of other points
in Fig. 5.2-5a are indicated. The block diagram is a simulation of the differential
equation
(5.2-2)
(5.2-3)
Then
q = y — bxv — b2v
(5.2-4)
q — ij — b-ii) — b2v
(5.2-5)
Collecting terms,
y — b2v — {bx + ab2)v = b0v — ay — by (5.2-6)
318 State Variables and Linear Continuous Systems
(b)
Fig. 5.2-5
Comparing Eq. 5.2-6 with Eq. 5.2-1, the requirements for b0, bx, and b2 are found by
equating like terms. They are
b2 = e
bx + ab 2 = d or bx = d — ae (5.2-7)
b0 = c
Time-Varying Systems
Example 5.2-4. Draw a block diagram for the time-varying system governed by the
differential equation y + a(t)y + b{t)y = v.
Using amplifiers with time-varying gains, this is simulated as shown in Fig. 5.2-6.
Fig. 5.2-6
Nonlinear Systems
The simulation diagram for this system is shown in Fig. 5.2-8a. This system can also
be represented as in Fig. 5.2-8b. The block indicated by ( )3 takes the place of the two
multiplier blocks. In this case the function to be generated is f(y) = y3.
y + ayy + by3 = v
This is a case where both a function generator and a multiplier must be used. The
simulation diagram is shown in Fig. 5.2-9.
Example 5.2-7. If the functional relationship between the input y and the output e of a
block is defined as e = d(y), draw a block diagram for the differential equation
y + d(y) + by = v
(a)
(b)
Fig. 5.2-8
e = d(y) ^
Ideal limiting
Fig. 5.2-10
Sec. 5.2 Simulation Diagrams 321
(a)
(b)
Fig. 5.2-11
Multivariable Systems
V\ + tyi + 2y2 = vx
y* + V\ + Vi — v2
Fig. 5.2-12
Fig. 5.2-13
The concept of the transfer function H(s) was introduced for single
input-single output fixed linear systems in Section 3.5. For systems which
have more than one input or output, transfer functions between various
Sec. 5.3 Transfer Function Matrices 323
Example 5.3-1. A system with two inputs and two outputs is governed by the set of
differential equations
iji + 52/! + 6 = +
yx Vi 3vi +
Av2 + 8y2
y2 + 2/2 = Vi + 2i)2 + 2v2
VM s
|Q->- Yi (s)
s+2
|Q -* Y2 (s)
/■
1K
V2(s) 4
s+3
Fig. 5.3-1
324 State Variables and Linear Continuous Systems
Example 5.3-2. A two input-two output system is described by the differential equations
Vi + y2 = v1 + v2
2/2 + Vi = v2
Hence
s 1
(S + 1)CS - 1) 5 + 1
H (s) =
-1 1
Js + 1)(5 — 1) S + 1_
It should be noted that the transfer functions 1 /(s + a) and s/(s + a), commonly
shown on diagrams of this sort, can be obtained from a single integrator. This is shown
in Fig. 5.3-3. An important point to be stressed here is related to the fact that the transfer
function diagram of Fig. 5.3-2 uses three integrators. A conclusion that the system is of
V(s)
V(s)
Fig. 5.3-3
Sec. 5.4 The Concept of State 325
third order is incorrect. The original differential equations require only two integrators
for simulation, as shown in Fig. 5.3-4. The simulation diagram, obtained from the
original differential equations, always shows the correct order of the system. However,
the transfer function diagram should be used with the physical problem always in view;
otherwise an incorrect conclusion about the order of the system may be made. The
Fig. 5.3-4
transfer function diagram often masks the physical properties of a system. This point is
covered at greater length in Section 5.7, when the controllability and observability of a
multivariable system are discussed.
These sets of states and assignment must satisfy the following consistency
conditions for a state-determined system:
(1) The admissible input functions must be such that, if v^t) and v2(t)
are admissible input functions, then
t Item 2 allows for the possibility of a time-varying state space, and item 3 involves the
concept of a final or terminal state for each initial state at time t. These ideas differ
somewhat from the more general ideas expressed by Zadeh and Desoer,1 whose concept
of state evolves from the characterization of a system by a listing of all observable input-
output pairs.
Sec. 5.4 The Concept of State 327
Fig. 5.4-2
and future inputs to a system uniquely determine the present and future
outputs. Thus, for every x(/0) contained in Xto and admissible input
functions Y^t), v2(0 with v^r) = v2(0 for t > t0, any output function
associated with x(70) and \±(t) is identical with any output function asso¬
ciated with x(>0) and v2(/) for t > t0. This is illustrated in Fig. 5.4-2 for
the scalar case.
(3) If the initial state of a system and the input v(0, t > t0, are given,
then the output y(t), t > t0, is uniquely determined. Assume that the
328 State Variables and Linear Continuous Systems
Fig. 5.4-3
known input and output time functions are divided into two time intervals.
For the second interval, there may be many initial states for which the
given input function over this interval yields the given output function.
However, at least one of these possible initial states must be the terminal
state of the first interval. This is illustrated in Fig. 5.4-3. More formally,
for any x(t0) contained in Xto and input function v(7), let Xx contained in
Xt , tx > t0, be the set of all states for which the output function associated
with \(t) and x contained in Xx be the same as the output function asso¬
ciated with v(7) and x(f0) for t > tx. Then x(A) is contained in Xx.
These three consistency conditions can be written as a pair of equations,
which are called the state equations. They are
where both g and f are single-valued functions. Equation 5.4-1 states that
the output y over the time interval t0 to Ms a single-valued function of the
state at the beginning of the interval and the input v over this time interval.
The state at the end of the interval is said, in Eq. 5.4-2, to be a single¬
valued function of the same argument. These two equations define a
state-determined system.
For most of what follows, the outputs of the integrators in the simulation
diagram are used as the components of the state vector. The state vector
is defined in terms of an //-dimensional state space, whose coordinates are
aq, aq,. . ., xn. The motion of the tip of the state vector in state space is
called the trajectory of the state vector.
Although the outputs of the integrators in the simulation diagram form
Sec. 5.5 Matrix Representation of Linear State Equations 329
where A(t), B(t), C(t), and D(t) are, in general, time-varying matrices, and
x± 01 Vi
x2 ^2 V2
• •
, V = » y =
* • •
• •
0m yf
A general diagram for these equations is shown in Fig. 5.5-1. If the system
is fixed, i.e., nontime-varying, then the matrices A(t), B(t), C(t), and D(0
are constant and can be written simply as A, B, C, and D.
Fig. 5.5-1
330 State Variables and Linear Continuous Systems
Example 5.5-1. Write a set of state equations for the system of Fig. 5.2-12.
'x, ro l 0 0- -Xg -o 0-
x0 o -3 -2 0 x2 1 0 Vi
+
x. o 0 0 1 x3 0 0 v2
X. o -1 -1 0 X 4_ _0 1_
x.
Vi '1 0 0 0' X, Vi
+ [0 0]
V2 0 0 10 X, v2
l_X4_
Thus
-o 1 0 0- “0 0-
0 -3 -2 0 1 0 T 0 0 O'
A = , B = C = D = [0]
0 0 0 1 0 0 0 0 1 0
_0 -1 -1 0_ _0 1_
Example 5.5-2. Determine a set of state equations for a general nontime-varying
linear network consisting of resistors, inductors, and capacitors.
Without loss of generality, the network can be represented as in Fig. 5.5-2a. Define
the state variables as
x4, x2,. . . , xm — voltages across the resistive network terminal
pairs connected to Cu C2,. . . , Cm, respectively; and
This value is the same as the constant current which would flow through the zth terminal
pair, if all capacitors were replaced by short circuits and all inductors replaced by open
circuits, and a source of 1 volt applied as abo.ve. Thus
Or ..
Clij
C*)
~Ci
i, j, 1,2, m
where is the current through the short circuit replacing Cu due to the unit step
voltage source replacing C3. Note that the resultant network which must be analyzed to
determine these au's is a purely resistive network.
Sec. 5.5 Matrix Representation of Linear State Equations 331
*771+1
Lm+ 1
Lm + 2
Ln
eL
For this same step of voltage applied in series with the y'th capacitor at t = 0, the
voltage across the /th inductor is
Thus
ru
@ij i, j= m + m + 2, . . . , n
Li
where ri} is the voltage across the /th open-circuited terminal pair in response to a unit
current source applied to the /th terminal pair.
For the same current applied in parallel with the y'th inductor at t = 0, the current
through the /th capacitor at t = 0+ is
where fin is the current through the /th short-circuited terminal pair in response to a unit
current applied to the /th terminal pair.
By these simple steps, the A matrix for a general time-invariant, linear RLC network
can be found by examination of purely resistive networks. As a specific illustration of
the procedure, the reader can confirm that the A matrix for the circuit of Fig. 5.5-2b is
2 0
A = (R\ + ^2 )C
2R±R2
0 ” CRi + Rz)L
y + ay + by = v
x2 = —bx i — ax2 + v
Fig. 5.5-3
Sec• 5-5 Matrix Representation of Linear State Equations 333
X, 0 1 ' x. 'O'
+
x, —b —a Xr 1
x,
y = [l 0] + [0]u
x*
Thus
‘ 0 1 ' O'
A = B = , C = [1 0], D = [0]
—b —a 1
y — x i + b0v
+ bkv, k < n (5.5-4)
= ~ (¥i + +*’*+ an-\xr) + bnv
Fig. 5.5-4
334 State Variables and Linear Continuous Systems
Following this procedure, the second and higher derivatives of y are given
by
Substituting for y, py, . . . ,pn~1y from Eqs. 5.5-4, 5.5-5, and 5.5-6 into
Eq. 5.5-3, and comparing the result with the expression forpny as given by
Eq. 5.5-6, the expressions for the at and bi are given by
at = oq (5.5-7)
and
b0 = Pn
b\ P &n—-\bo
^2 = Pn-2 an-\bl ~ arz-2^0 (5.5-8)
1 0 0 • • • 0
o
Pn
P n—1 <*„_ 1 1 o • • • 0 bi
• = • • • • • •
• • • • • •
.
• • • • • •
.
0 1 0 • • •
0
0 0 1 • • •
0
A = (5.5-9)
• • • • • • • • • • • •
1
— — a2 • • •
— a0 ai
b0 1 0 0 O' ftn
ftn-1
bi an-1 1 0 0
ft n—2
b2 ^rc—2 ^n—1 1 0
D
=
• • •
B
• • •
• • •
• • • a n—1 ftc
_ ^0 ai
• • •
[10 0 0]
o
II
o
Pn
y + 3i) + Ay + y = 2v + 3v + v + 2v
in standard form.
From Eq. 5.5-8,
b0 = fin = 2
bi = fin—i - v-n-ibo = 3 — 3(2) = -3
b 2 = fin-2 — *n-ibi — a n_2b0 = 1 — 3(—3) — 4(2) = 2
bo = fin-2 — *n-ib2 — an_2^i — a„_360 = 2 — 3(2) — 4(—3) — 1(2) = 6
Thus Eqs. 5.5-2 and 5.5-9 give as a vector matrix representation for this system
Xi
y = [1 0 0] x2 + 2y
xo
Fig. 5.5-5
and
N(s)
Ci = (s - A,.)
D(s) s=A,
n c
Y(s) = d0V(s) + 2 —s- K(s) (5.5-11)
s Aj 2=i —
Sec. 5.5 Matrix Representation of Linear State Equations 337
*i - Kxi = v
and
n
y = 2 cixi + do V
2=1
x-, «1 1
1
o
Xc
0 • • • 1
0 22 • • • 0 •
+ V
•
.
•
0 0 • • •
1_
s
Xn _Xn_ 1
(5.5-12)
Xi
Xo
Xn
Fig. 5.5-6
338 State Variables and Linear Continuous Systems
or
x = Ax + vl
(5,5-13)
y = (c, x) + d0v
For the case where the roots of the denominator polynomial D{s) are
not distinct, the partial fraction technique leads to a nondiagonal Jordan
canonical form. As an example of this, consider the case where D(s) is of
the form
D(s) = dn(s - Xy(s - A2)(j - A3) • • • 0 — K)
cn 12 Ik
H(s) — d0 + + k-1 + * • • +
(s - Kf (* - 0 - ^i)
Co Co e„
+ -+-2— + • • • +
S — 2o S /i-: s — Xn
n
<T3- C„■
— d0 + 2 +
f1 (s - xyk~i+l) ' i“* s - A,
2 (5.5-14)
where
d0 = lim H(s)
S-> 00
1 v-1 kN(s)
clA = (s - AJ
U - 1)! ds*-1 D(s)J s=;.x
and
N(s)
Ci = (s - A,)
D(s) s=A,
1
n(s) = n+i(s) 7 = 1.fc — 1 (5.5-16)
(s - Aj)
Sec. 5.5 Matrix Representation of Linear State Equations 339
0
Xr ** + 1 v
Xk+1
0 0 0 Aj 0 0 0
%-t-l
1
0 0 0 0 A2 0 0 • •
•
• • • • • • •
• •
0 0 0 0 0 0
Xn _*n _ 1
(5.5-17)
340 State Variables and Linear Continuous Systems
Xi
x2
Xr
y — [cn ci2 -Ik C«,
cj + d0v
Xk+1
or
x = Jx + bv
(5.5-18)
y = <c, x) + d0v
where J, b, and c are defined by the equivalence of Eqs. 5.5-17 and 5.5-18.
The partial fraction technique is appealing when the system has only one
input and one output. The result of this approach is an A matrix which is
either in diagonal form or a nondiagonal Jordan form. This form is
particularly convenient when the dynamic behavior of the system is to be
found. However, this approach presents certain difficulties. Most of
these difficulties can be overcome by the general technique of converting
the state equations in standard form (if they are available) to the normal
form.
Normal Form
The method of partial fraction expansion is useful when the state equa¬
tions of a single input-single output system are to be found. In this case
B, and C are vectors and D, v, and y are scalars. When the system is
multivariable, the method of partial fraction expansion can become un¬
wieldy. It then becomes desirable to be able to convert the state equations
in standard form to a form where the A matrix is either a diagonal matrix
or a more general Jordan matrix. This conversion can be accomplished
by means of a similarity transformation.
Consider the fixed system defined by the state equations in the standard
form
x = Ax + Bv
(5.5-19)
y = Cx + Dv
x +
y = [1 0]x + 2v
The modal matrices M and M-1 associated with the A matrix are
" 1 1 " 2 1
M = and M_1 =
-1 —2 _ -1 -1_
342 State Variables and Linear Continuous Systems
(a)
(b)
(c)
Fig. 5.5-8
A = M XAM =
2
B = M B = n
Cn = CM = [1
Dn — d o — 2
Sec. 5.5 Matrix Representation of Linear State Equations 343
q = Aq + bnv
y = <cq> + d0v
where q = M_1x, or qx = 2a?! + x2, q2 = —xx — x2. The block diagram for these
equations is shown in Fig. 5.5-86.
The transfer function of this single input-single output system is
2s2 + Ss + 11 2s2 + Ss T 11
H(s) =
s2 + 3s + 2 (5 + 1)(^ + 2)
The partial fraction expansion of H(s) is then
d0 = lim H(s) = 2
s—► 00
c, = H(s)(s + Dl,..! = 5
c2 = H(s)(s + 2)|»__2 = —3
Consequently
5 3
H(s) = 2 +
s + 1 s + 2
The state equations of this system are
"-1 0“ ~r
X = x +
0 -2 1
y — [5 — 3]x + 2v
Example 5.5-6. Write the state equation for the system of Fig. 5.2-13 in normal form.
For the state variables indicated in Fig. 5.2-13, xz is already in normal form. Thus
only xx and x2 need be considered. Then
" 0 r xx "0 v1
+
-2 -3 x2 1 v2
pi
y 2 = [i 0]
L^2_
or
c = [1 0], D = [0]
“ 1 r 2 r "-1 O'
M = , M-1 = , A =
-1 — 2_ -1 -1 0 -2_
344 State Variables and Linear Continuous Systems
Hence
B = MB = Cn = CM = [1 1]
F(s)
2/(0 = se -l
+ [22/(0) + 2/(0)]€-‘ - [2/(0) + 2/(0)]<T2'
_fs T- l)fs + 2T
Sec. 5.6 Mode Interpretation—A Geometrical Concept 345
If the terms of the response involving <rl and e~2t are called “modes” of
the system, then the solution y(t) may or may not exhibit these modes,
depending upon the initial conditions and the zeros of F(s). For example,
if the initial conditions are such that 2y(0) + 2/(0) = 0, then the initial
condition response does not exhibit the mode If, in addition, F(s)
has a zero at s = — 1, then the complete solution for y(t) does not exhibit
the mode The mode e~l is completely suppressed. Consequently, if
the initial conditions are such that a zero of the response transform occurs
at the same point as does a pole of the transfer function of the system, and
if a zero of the forcing function transform also occurs there, then the mode
corresponding to that pole does not appear in the system response. This
pole-zero cancellation, although somewhat obvious, leads to the general
concepts which are to be discussed.
Since the characteristic roots are distinct, the characteristic vectors are
linearly independent. Therefore x(t) can be uniquely expressed as a linear
combination of these characteristic vectors:
n
By forming the scalar product of both sides of Eq. 5.6-6 with the con¬
stant ci is found to be
c, = <r„ x(0)> (5.6-7)
The scalar product (ri? x(0)) represents the magnitude of excitation of the
ith mode of the system due to the initial conditions. If the initial conditions
initially lie along the zth characteristic vector, then only the ith mode is
excited. The scalar products (r3, x(0)), where i ^ j, are identically zero.
Therefore, for a linear, fixed, unforced system with distinct characteristic
values, the initial condition response is given by a linear weighted sum of
the modes where is a characteristic root of the A matrix.
Example 5.6-1. To illustrate the use of the mode expansion technique analyze the
system considered at the beginning of this section.
For this system and dC OC ^ j OC “ OC 2 j
1/V5'
u2 =
-2/V5
and
1 <N
<
1_
>
1
<N
1
_1
j r2 —
1
>
1
_
1
x(o = 2 eXitvLi
i=1
Note that, if the initial condition vector is set equal to one of the characteristic vectors
(or any factor times the characteristic vector), then the scalar product (r*, x(0)> vanishes
for all but the component associated with that characteristic vector. For example, if the
initial conditions are such that y{0) = 1,2/(0) = — 1, then the initial condition vector
lies along ux and only the mode involving is excited.
Sec. 5.6 Mode Interpretation—A Geometrical Concept 347
So far, only the homogeneous case case been discussed. For the case in
which a forcing function is applied to the system, the same general prin¬
ciples apply if the forcing function is first decomposed among the
vectors as
n
Bv = f(f) = 2//0U, (5.6-9)
i=1
where ft(t) = (r*, f(t)> = (ri? Bv(t)). Utilizing the convolution integral for
linear fixed systems, the general expression for x(t), assuming distinct
characteristic values, is then given by
The importance of Eq. 5-6.10 lies in that the effect of the forcing function
on each mode is considered independently. The amount of excitation of the
zth mode due to the forcing function is given by
| Vi, Bv(r))e'l<(<-'>ui dr
Jo
If the forcing function is so selected that it always lies along the direction
of one of the characteristic vectors, then only one mode is excited by the
forcing function. This situation occurs in circuits that have symmetrical
or balanced properties. In these circuits, the modes are referred to as
symmetrical or antisymmetrical, such that a “symmetrical” forcing func¬
tion excites only the symmetrical modes.3
The characteristic vectors are found by successively substituting the A/s into Adj [Al -A].
Since
"A + 2 1
Adj [AI - A] =
-2 A
r/ = V/- 0
6 V6
T
The complete solution is then
If the initial conditions are so chosen that x(0) = u/, then only the mode
is excited. If the initial conditions are so chosen that x(0) = u/', then the mode
is excited. The difference between these two modes is the amplitude and phase of the
damped oscillations.
x = Ax + Bv
(5.7-1)
y = Cx + Dv
350 State Variables and Linear Continuous Systems
then the contribution to x(t) due to the forcing function can be expressed
as
t 72 LYl
2 2 <ri> bJ.)uXT)€'iilt“,)Ui dr
Jo 1=1 0=1
If the scalar product (iy, b;> is zero for some mode for ally, then the input
is not coupled to that mode, and, regardless of what the forcing function is,
there is no way for the input to excite or control this mode. Obviously,
a criterion for complete controllability of a linear time-invariant system
is that the scalar products (ly, b3-) do not vanish for ally.f
A system which is not observable has dynamic modes of behavior which
cannot be ascertained from measurements of the available outputs.8
Considering only the initial condition excitation of the modes, the output y
is given by Eqs. 5.7-1 and 5.7-2 as
n
assuming that A has distinct 2’s. A condition that the ith mode disappear
in all the outputs is
(c u,-) = 0 j,for all j (5.7-5)
t This is a necessary and sufficient condition if A has distinct A’s. If A has repeated
/Ts and is diagonalizable, then this is necessary but not sufficient. In this case or if A
can only be transformed to a more general Jordan form, then one should consider the
more basic requirement for controllability, which is that there be an input \(t),
0 < t < T < oo, such that x(0) can be forced to x(T) = 0. This leads to the necessary
and sufficient condition that the column vectors of the matrix [B, AB, . . . , A(n-1)B]
span the state space of the system.
Sec. 5.7 Controllability and Observability 351
t If A has repeated A’s, the basic requirement for observability is that, for some
T > 0 and all initial states x(0), knowledge of A, C and y(t), 0 < t < T, suffices to
determine x(0). This leads to the necessary and sufficient condition for observability
that the columns of [C*r, A*TC*T, . . . , (A*r)(n-1,C*7'] span the system state space.
352 State Variables and Linear Continuous Systems
Fig. 5.7-1
Composite Systems
When a system is comprised of some or all of the components S*, Sc, S°,
or Sf, the overall system may or may not exhibit the properties of these
subsystems, depending upon the connections between the subsystems.
Example 5.7-1.
Vi = X1 +
Now let the output of system 1 be connected to the input of system 2, and let the output
of the system be taken as y = x2 — yx. This is shown in Fig. 5.7-2. Determine the
controllability and observability characteristics of the composite system.
Fig. 5.7-2
Sec. 5.7 Controllability and Observability 353
x1
+
x2
y = [—l Vi
1/V2
u, =
_1/V2_
1
r2 =
1
For this system,
(r„b> = V 2 <r2, b) = 0
<c, ux> = 0 (c, u2> = -1
Therefore the mode e~2tu2 is not controllable, and the mode is not observable.
Thus, even though the individual subsystems are both observable and controllable, the
overall system is neither observable nor controllable.
An interesting point arises if transfer functions are utilized. The transfer function of
the first system is (s 4- 2 )/(s + 1). The transfer function of the second system is l/(s + 2).
Therefore the overall transfer function from vx to x2 is
X2(s) _ 1
V2{s) s + 1
This confirms the previous analysis that the mode e~2'u2 cannot be excited and the mode
cannot be observed. This analysis by means of transfer functions shows that the
use of a transfer function for an overall system may mask some of the modes of the
subsystems.
(a) The order n of the system is equal to the sum of the orders of Sa and
Sb, i.e., n = na + nb.
(b) A necessary and sufficient condition that the feedback system be
controllable (observable) is that SC(S0) be controllable (observable).
354 State Variables and Linear Continuous Systems
ya = y
yb V6
sb
Fig. 5.7-3
(c) A necessary but not sufficient condition that the feedback system be
controllable (observable) is that both Sa and Sb be controllable (observable).
(<d) If Sa and Sb are both controllable (observable), any uncontrollable
(unobservable) coordinates of the feedback system are uncontrollable
(unobservable) coordinates of SC(S0) and originate in Sb.
The importance of this theorem lies in the fact that controllability and
observability can be determined from the individual open-loop subsystems
Sa and Sb. This is a great aid in analysis.
Transfer Functions
t If A has repeated characteristic values, this may not even be true. For example,
consider such a case in which A is diagonalizable.
Sec. 5.7 Controllability and Observability 355
Example 5.7-2. Determine the transfer function matrix and the order of the system
described by
Vi + ayx = vx + v2
y-z + {a + b)y2 + aby2 = vx + avx + v2 a ^ b
Vx(s) V2(s)
4
Y1(s) = - -i +
s + a s + a
Vx(s) V2(s)
Y2(s) = -^-r +
s + b (s + a)(s + b)
The transfer function matrix is
■ 1 1 “1 1 " “0 0 ”
5 + a s + a 1 1
H0) = = 0 7- + 1
1 1 b — a a — b
_s + b (s + a)(s + b)_ s + a s + b
The immediate, but erroneous, conclusion is that this matrix represents a second
order system. However, if a simulation of this system is made, the minimum number of
integrators required is three. This is shown in Fig. 5.7-4. This fact is evident from the
original differential equations.
H(s) = 2 + D t5'7’7)
i—l S — At-
where
K- = lim (s — A*)H(s) and D = lim H(s)
S-+A,- s-» oo
It is assumed that the elements of H(s) have, at most, m finite simple poles,
where m < oo. The rank of the zth pole, ri9 is defined as the rank of the
356 State Variables and Linear Continuous Systems
n = Xri
i=1
The matrix 4>(/) = eAf is called the state transition matrix of the system
described by Eq. 5.8-1. Mathematicians prefer the term fundamental
matrix.% The nomenclature state transition matrix is more descriptive of
t For complete discussion of this problem, including the case where the characteristic
values are not distinct, the reader is referred to References 5, 8, and 10.
± For a rigorous discussion of fundamental matrices and their role in the solution of
differential equations, see Reference 11 or 12.
Sec. 5.8 Linear Fixed Systems—The State Transition Matrix 357
From the definition of eAt (Eq. 5.8-3), the state transition matrix 4>(7)
can be calculated by the infinite series
\2t2 A 3t3
<K0 = I + A? + — + — + • • •
2! 3!
Unless disappears for some small value of k, this method is the most
laborious. Once the summation is performed, the closed form of each
series for each element of the<J>(0 matrix must be found. This is generally
not a simple task, and, unless the order of cj>(?) is sufficiently low or the
analyst is sufficiently clever, the task may be insurmountable. A simple
example points out the difficulty involved.
Example 5.8-1. Find the state transition matrix <f>(0 for the matrix
'0 - 2'
A =
1 -3
-2 6 '6 14'
A2 = and A3 =
-3 7 7 -15
_0 1_ _1 —3_ -3
i
358 State Variables and Linear Continuous Systems
By recognizing the infinite series for each element (this is the principal drawback of this
method),
r2e-J - e~2t 2(e~2t 6-01
<*>(') =
2e~2t
Frequency-Domain Method
If the Laplace transform of both sides of Eq. 5.8-1 is taken, the result is
sX(s) - x(0) = AX(j). Thus
This method may be the most convenient to use for many problems. The
obvious difficulty is finding the inverse of [si — A].
Example 5.8-2. For the same A matrix as in the previous example, calculate <J>(0
using the frequency-domain approach.
Since
^ 2
[si ~ A]
-1 5 + 3
where <E>(s) is the transform of <}>(0- T aking the inverse transform of <h(s) element by
element, the expression for is found to be the same as previously computed in
Example 5.8-1.
Sec. 5.8 Linear Fixed Systems—The State Transition Matrix 359
The zth term x^t) of Eq. 5.8-2 can be written as the summation
Example 5.8-3. The simulation diagram for the system described by the A matrix used
for the previous two examples is shown in Fig. 5.8-la. Determine
(a)
With no loss in generality, the system can be redrawn in the frequency-domain form
shown in Fig. 5.8-16. All the transfer functions are then immediately evident.
Vs s + 3
#u0)
2 s2 + 3s + 2
s(s + 3)
-2
s(s + 3) __ -2
#12 (S)
“ 2 s2 + 3s + 2
+ s(s + 3)
1
s(s + 3) 1
#2lO)
2 52 4- 3s + 2
+ s(s + 3)
1
s + 3 _ s
#2 2Cf)
2 s2 + 3s + 2
s(s + 3)
Hence
s + 3 -2
1 s_
<£(s) =
s2 + 3s + 2
This matrix checks with ^(s) found in Example 5.8-2 by computation of [si — A]-1.
The calculation of <J>(7) is then simply the inverse transform of ^(s), element by element.
The complete set of state variable equations for a linear fixed system is
given by
x = Ax + By
(5.8-8)
y = Cx + Dv
However, the first term is a zero matrix, since cj>(7 — r) is a matrix whose
columns are solutions of Eq. 5.8-1. Thus
Finally, since
_ A(<—t) — A(A—t)
<(>(0) = I and <\>(t — r)4> (2 — t) = € 7,e
362 State Variables and Linear Continuous Systems
The solution for y{t) follows by substituting Eq. 5.8-11 into Eq. 5.8-8. It is
Equations 5.8-11 and 5.8-12 form the solutions to Eq. 5.8-8. The first term
of Eq. 5.8-11 represents the initial condition response of the system state
variables, while the second term represents the forced response. Note that
the second term of Eq. 5.8-11 is a convolution integral similar to that in
Eq. 1.6-2.
Example 5.8-4. Assume that the system whose A matrix is given by Example 5.8-1 is
subject to a unit step forcing function at t — 0 (see Fig. 5.8-la). Find the output
2/(0 = «i(0-
For this example,
'0 —2' '2e-* - e--2i 2(e~2t - e-*)'
A = , <*>(0 =
1 -3 €,-t' - e-2t 2e~2t -
O'
B = , C = [l 0], D = [0]
1
Therefore, from Eq. 5.8-12,
-1
dX
©
+
= [i 0]
_x2(0)_ Jo _ 1_
For the case where the A matrix is not fixed but varies with time, the
homogeneous matrix differential equation of a linear system is
x = A(t)x (5.9-1)
Sec. 5.9 Linear Time-Varying Systems—The State Transition Matrix 363
x = a(t)x (5.9-2)
considered in Section 2.5, and the solution to Eq. 5.9-1. The solution to
Eq. 5.9-2 is
x(t) = [exp b(t)]x(t) (5.9-3)
b(t) a(X) dX
However, if Eq. 5.9-4 is substituted into Eq. 5.9-1, it is seen that Eq. 5.9-4
is the correct solution if and only if
B(0 = A(2) dX
If Eq. 5.9-6 is satisfied, then Eq. 5.9-4 is the solution to Eq. 5.9-1 and the
state transition matrix is given by
The Matrizant
For the general case, where the commutativity condition of Eq. 5.9-6 is
not satisfied, the state transition matrix is not given by Eq. 5.9-7. However,
364 State Variables and Linear Continuous Systems
the solution to Eq. 5.9-1 can be obtained by a method known as the Peano-
Baker method of integration.14 This method follows.
Let the conditions x(r) be given. Then, by integrating Eq. 5.9-1, the
integral equation
If the process shown in Eq. 5.9-9 is continued, then x(t) is obtained as the
Neumann series
The first term in the parentheses is I, the unit matrix. The second term is
the integral of A between the limits r and t. The third term is found by
premultiplying Q(A) by A and then integrating this product between the
limits r and t. The other terms are found in like manner. If the elements
of A remain bounded between the limits of integration, then this series is
absolutely and uniformly convergent. This series defines a square matrix
G(A) which is called the matrizant.
If both sides of Eq. 5.9-12 are differentiated with respect to t, the funda¬
mental property of the matrizant
is obtained. Therefore G(A) is indeed the solution to Eq. 5.9-1 and, as such
represents the desired state transition matrix for a time-varying system.
Thus
4>(*. r) = G(A) (5.9-14)
or
x(r) = G(A)x(t) = 4>(r, t)x(t) (5.9-15)
it2 - T
G(A) = <fi(t, r) = 1 - —
Recognizing this as the form for the infinite series e-2, the state transition matrix is then
given by
<Kt, r) = exp
This simple example points out the difficulty in using the matrizant
approach. Unless the series (Eq. 5.9-12) converges rapidly, the com¬
putation becomes quite lengthy.
An interesting alternative solution was proposed by Kinariwala.13 The
approach is to decompose A(t) into two matrices, A0(t) and Ax(t), where
A0(f) satisfies the commutant condition of Eq. 5.9-6. Thus
The solution to Eq. 5.9-17 is assumed to be of the same form as Eq. 5.9-19,
but with successive corrections added to take the perturbations into ac¬
count.
The perturbations Ai(0x(0 are equivalent to the forcing function term
Bv of the previous section. By direct analogy, using the superposition
theorem, the solution for x(t) is then given by
rt
x(0 = 4>o(*> t)x(t) + <J>o0, ^)A1(A)x(A) dX (5.9-20)
J T
Again, this is a Yolterra equation of the same form as Eq. 5.9-8. Using
the same process of iteration as previously described, the Neumann series
for<|>(/, r) is found to be
Since
z‘i — + tz2
dt
Sec. 5.9 Linear Time-Varying Systems—The State Transition Matrix 367
(a)
(b)
Fig. 5.9-1
then 22 = ?! + tz2. The expression for zx = y can be obtained from the definition of z2.
Thus zx — y = z2 -f- tzx. The matrix differential equation is then
«1 t r
2_ _1 t_ _22_
The simulation diagram is shown in Fig. 5.9-1 b. The new A(t) matrix
t r
A (t) =
1 t
does obey the commutant condition, i.e., A(tx)A(t2) = A(t2)A(tx). Therefore Eq. 5.9-7
can be used.
The state transition matrix 4>(r, r) is given by
where
t T
The matrix eB can be found by any one of the techniques previously discussed. However,
it is instructive to find eB by use of Eq. 4.10-20. Thus
eB = MeAM 1
368 State Variables and Linear Continuous Systems
The characteristic values of B are Ali2 = [(72 — r2)/2] ± (/ — r), and the modal matrices
are
-1 r "1 r
M = (/ — T) M_1 — ^
1 -i_ 2(7 - r) _1 -i_
Therefore
t2 - T2 0
i r exp + (7 “ r) T T
72 - T2
i -l
exp — (r — r) 1 -1
0
and
The matrix <j>(7, r) does satisfy the differential equation (dldt)[<\>(t, r)] = A(r)4>(7, t),
since
'0 1
-j 4>(T r) = *4>(T t) + <t>(7>T)
dt 1 0
and
7 1‘ ‘o r
A(/)4>(/, r) = <t>0, T) = t) + 4>0, t)
1 r 1 o
Although almost all the time-varying problems that the control engineer faces must be
solved by either an analog or digital computer, it is also true that careful prior inspection
of the system can greatly reduce the computation time. The obvious state variables may
not be the state variables that should be used to control the system, or used to simulate
the system. In many cases, combinations of the obvious state variables may be more
valuable. The foregoing example illustrates this point.
The state transition matrix can also be found by using the simulation
diagram method of Section 5.8. For a time-varying system, the ith state
variable xj(t) is given by Eq. 5.9-22, similarly to Eq. 5.8-7.
The term </>^(/, t) can be found by obtaining the transfer function O^(t, s)
between the input to integrator j and the output of integrator /. Note that
this is a time-varying transfer function (see Section 3.9) and as such may be
difficult to evaluate. If the system were actually simulated on an analog
computer, a unit initial condition placed on integrator j at time r1 would
produce the response <f>i:j(t, tj) at the output of integrator /. A series of runs
and a cross plotting of the results are necessary if r) for all r < r is to
be obtained. This point is discussed in detail in Sections 5.11 and 5.12.
Sec. 5.9 Linear Time-Varying Systems—The State Transition Matrix 369
Property 1: By definition,
4»Oo, to) = I (5.9-23)
This can be shown from the relations x(r2) = t1)x(t1) = 4>(/2, toMto)
and x(0 = <!>(/!, f„)xOo)- Then x(t2) = <f>(/2,0+Oi, foM'o)- Therefore
<&(h, t0) = 4>fo h)4>(/i, t0)-
Property 3:
4>(h,t2) =4>-1(t2,t1) (5.9-25)
<K*, r) = 4r*(T, 0
The behavior of the system with respect to the variable t is a function of the
dynamics of the original system. The behavior of the system with respect
to the second variable t is a function of the dynamics of the system for
which 4>-1(?, r) is the state transition matrix. This system is known as the
adjoint to the original system.
370 State Variables and Linear Continuous Systems
where a is a column vector. This can be derived from Eq. 4.4-8, which
indicates that
Therefore <J>_1(/, r) is the state transition matrix for the system whose
unforced differential equation is given by
a = — aA (t)
where a is a row vector. If the transposes of both sides of Eq. 5.9-30 are
taken, then|
Thus [4>T(/, t)]_1 is the state transition matrix for the system whose un¬
forced differential equation is given by
a = -AT(0<x (5.9-31)
where a is a column vector.
Example 5.9-3. Compare the state transition matrices of x -f tx = 0 and the adjoint
equation.
From Example 5.9-1,
or
/r2 - t2\
<Kt, r) = exp I —-— I t > t
= 0 t < T
or
It2 - r2\
r) = exp I —-— I t > t
= 0 t < T
/r2 - t2\
t) = exp I—-— I T > t
= 0 r <C t
A comparison of , t) with <f>{t, r) shows that the two expressions are identical
except that they are valid over different intervals. The physically realizable output
a(t) of the adjoint system represents the physically unrealizable portion of the solution
x(t). A physically realizable output of a system implies that the observation variable
is greater than the application variable.
= pn + y ak{t)pk, pk = J-
k—0 at
and the ak(t) are real. The linear adjoint differential operator is defined as
wherepkan(t) signifies that pk operates on the product of ak{t) and the depen¬
dent variable. Consequently, the linear adjoint differential equation
Ln * a = 0 can be written as
T 0 1 0 0
0 0 1 0
A = (5.9-34)
0 0 0 1
-a0(t) —#i(0 —a2(t) • * ‘ ~an-i(t)_
372 State Variables and Linear Continuous Systems
0 0 • • • 0 floO)
-1 0 • • • 0 «i(0
0 -1 • • • 0 a2(t) (5.9-35)
_ 0 0 • • • -1
To show that the operator notation and the matrix notation are equivalent,
the following observations are made on the adjoint matrix formulation.
«i 0o(O<*n
a2 — <*i + a1(t)ccn
ax ~ a<- 1 +
Differentiating an,
= a«-2
dt
If this process of differentiation and substitution is performed (« — 1)
times, it can be shown that
matrix C such that <J>(/ + T, r) — <f>(t, r)C. This can be shown rather
easily, since the derivative of this expression yields
P(0 is nonsingular, since <J>(/, t) and are nonsingular for — oo < t <
oo. Therefore, for a periodic system, P(t) is a nonsingular periodic matrix
with period T.
It is interesting to note that P(/) is the solution to the differential equa¬
tion17
P(0 = A(/)P(0 ~ P(0» (5.9-39)
The term p{t) is periodic, with period T, if b is equal to the average value of apt) over an
interval T, i.e.,
b =
i
-
rTa(X) dX
1 Jo
Any integral multiple of l^j may be added to this value.
y + oj\t)y = 0
where the periodic coefficient eo2(0 is shown in Fig. 5.9-2. Note that it is piecewise
constant. The A matrix for this equation is given by (xx — y, x± = x2)
o r
A =
-0j\t) 0
1 .
cos cot —sin cot
<f> = _ A]-1 co
—oo sinojt coscot
oo2(t)
Fig. 5.9-2
Sec. 5.10 Linear Time-Varying Systems—The Complete Solution 375
For the interval —tt/2 to tt/2, oj = V2 for a2 — b2 = 1, and co = 3.1 for a2 = 4.6,
b2 = 5.0. Thus
II
\2'-2) = -1.37 —0.259_ 0.959 —0.951_
IN
>
3
II
For the interval tt/2 to 3tt/2, oj = 0 for a2 = 62 = 1, and co = j0.634 for a2 = 4.6,
b2 = 5.0. Thus
/ 37T A - ”i 77 "3.73 5.66”
U ’V ~
_0 1 a>=0
2.21 3.73_ co=J0.634
Using the results of the previous section, the complete solution to the
state equations for a linear time-varying system can be obtained. The
state equations for a linear time-varying system are given by
x = A(/)x + B(/)v
(5.10-1)
y = C(f)x + D(1)V
Equation 5.9-30 states that
4>_1(b t) = t)A(/)
Since <j>_1(r, r) = I,
The expression for y(t) is obtained by substituting Eq. 5.10-6 into the
second of Eqs. 5.10-1, to yield
Equations 5.10-6 and 5.10-7 represent the complete solution to Eq.5. 10-1.
These equations are quite similar to Eqs. 5.8-11 and 5.8-12, respectively.
For a fixed system, the response depends upon t — r, or the time
difference between application of cause and observation of effect. The
result is the convolution integrals of Eqs. 5.8-11 and 5.8-12. The state
transition matrix for a fixed system has only one variable, namely the time
difference between application of cause and observation of effect. How¬
ever, for a time-varying system, the solution depends upon both t and r.
For a time-varying system, the cause-effect relationship is varying with
time, and therefore when the cause is applied and when the effect is ob¬
served are significant. The state transition matrix for a time-varying
system has two variables, one being the time of application of the cause,
and the other being the time of observation of the effect. Therefore
Eqs. 5.10-6 and 5.10-7 depend upon superposition integrals and not upon
convolution integrals.
Jrdi dX = dX (5.10-9)
f^a(a) da
€ —
X(t) — € —STa(<x) da x(t) = a(a) da
v(X) dX
-V
Thus x(t) can be written as
t n t
x(t) = €ST aU) docx(r) + a(a) dav(X) dX (5.10-10)
*) T
Notice the similarity between Eqs. 5.10-10 and 5.10-6. c(>_1, the state
transition matrix for the adjoint system, is the integrating factor used for
the solution of Eq. 5.10-1. This is a fundamental property of the state
transition matrix for the adjoint system.
A(0 =
Determine <J>0> T)-
A(t) does not satisfy the commutativity condition, since
12 "-I
Therefore the state transition matrix r) is not given by Eq. 5.9-7. The state transition
matrix could be obtained by use of the matrizant, but the series solution obtained by
this method is difficult to express in closed form. However, by use of Eq. 5.9-22, a
set of differential equations can be obtained which are solvable, if the integrating factor
approach is used.
Integrator Integrator
2 1
Fig. 5.10-1
378 State Variables and Linear Continuous Systems
y + ~r (yt) = 0
dt
For the requirements of this example, the expression is best left in this form. The
integrated form is quite complicated, as evidenced by
This illustrates an interesting point. Even when an analytical solution can be obtained
for a time-varying differential equation, this solution may not be easily evaluated. In
general, the solution for a time-varying differential equation is obtained by a computer
simulation, as the analytical solution is either difficult to obtain or difficult to evaluate.
The derivative y{t) is
The solution for the state transition matrix can now be obtained by application of Eq.
5.9-22. If a unit initial condition is placed on integrator 2, and zero initial conditions
on all other integrators, then the time response of the output of integrator 1, y(t), is
equal to <f>l2(t, r). Therefore let y(r) = 1 and y(r) = 0.
The expression for y(t) gives cx = 1. Therefore
j ft
t) €~t2/ 2 V2/2 dl = e(A2-i2)/2dX
* T
In a similar manner, the other terms of the state transition matrix are obtained from the
following:
<M', T) = V(0 when y(r) = 0, y(r) = 1
r) = y(t) when y(f) = 0, y(r) = 1
r) = y(t) when y(r) = 1 ,y(r) = 0
Thus
r) = + T dX
e(t*-«>)/2 + t £U>-(>)/2 M
j’Zl('. r) = T — t
t) = 1 — t\ €*A“ ^l*dX
Sec. 5.10 Linear Time-Varying Systems—The Complete Solution 379
e(rS-<*)/2 + TJ«W‘-<’,/a<W
4>(h t) =
r — t(f>n(t, r) 1 — rfls(t, r) _
The adjoint system is useful for studying the effects of forcing functions
and initial conditions on linear combinations of the state variables. Con¬
sider an inner product of the adjoint variables a and Lx, where
L = - A
(5.10-13)
from Eq. 5.9-29. Then, from Eqs. 5.10-11, 5.10-12, and 5.10-1,
Example 5.10-2. A terminal control system described by x = —(1 /T)x + m(t) starts at
ai(0) = 0 in response to m{t) = KU^{t). The state of the system at time t = t0 (a
fixed value) is desired to be #0(70) = €iti~to)IT, at which point m{t) is set equal to zero.
The system is then to “coast” to x = 1 at t = tx (also a fixed time instant). This is the
desired terminal point. When x(t0) is compared with x°(t0), however, it is found that,
owing to disturbances, x(t0) < #°(r0). How much longer should m(t) = K be maintained
so that x(tj) = 1 ?
Let At be the time duration beyond t0 for which m(t) = K. Then Eq. 5.10-14 becomes
't0+At
— <x(t0)x(t0) + I JCcc(X) dX
Comparison of Eqs. 5.11-1 and 5.11-3 shows that the impulse response
matrix H(f, r) is given by
Example 5.11-1. Find the impulse response matrix of the RLC circuit shown in Fig.
5.5-2b. The inputs are v^t) and v2(t), and the outputs are zffi) and z2(/).
382 State Variables and Linear Continuous Systems
o
(Ri + *2)C
A =
2
0 (Ri + R2)LJ
Since A is diagonal, <f> = eA< is easily determined.
Owing to the symmetry of the circuit, the relationships between the elements of the
coupling matrices are:
C11 = C21 («! has the same effect upon ix and i2)
c22 = — c12 (x2 has opposite effects upon i\ and i2)
du = d22 (vx couples into ix in the same manner as r2 couples into z2)
d\2 = d21 {vx couples into i2 in the same manner as r2 couples into
cn C 22 ^n(0 o bn bn dn d12
H(t) = + tuo
cn C22 0 (f>22 (t)_ b22 b22 d\2 d\\
1 1 -1 r dn di2
— ^llCll^ll(0 + b 22c 22(f> 22{t) + U0(t)
1 1 1 -l dn dn
From the state equations
x = Ax + Bv
y = Cx + Dv
for all t. Assume that the network is initially at rest (xx = x2 = yx = v2 — 0). If vx is
suddenly made 1 volt by application of a unit step of voltage, XjfO-l-) = ^2(0+) =
y2(0+) — 0, since the capacitor voltage and inductor current cannot change instan¬
taneously. Thus
4(0+)
bn= -
C
Vi=U_x(0
This is the same as replacing the capacitor by a short circuit and the inductor by an open
circuit (the network is now a purely resistive network) and writing
bn =
4(0 = _ gey
Cvi(t) C
where gcl is the short-circuit transfer conductance between the input vx and the short-
circuited capacitor. Thus
1
bn = -
(/+ + R2)C
Sec. 5.11 Impulse Response Matrices 383
In a similar fashion,
b = - 6l^ ^
22 Lv2(t) L
where ai2 is the voltage gain between the input v2 and the open-circuited terminals of the
inductor. The capacitor and vx are shorted. Therefore
R>
bn —
(Ri + R2)L
Using similar arguments,
h 1
cn — —, *2 = v, = v2 = 0 or cn =
Xi Ri + R2
Is R*
C 22 — , *1 = vx = v2 = 0 or c22
X« Ri + Ri
h 1
d1 i = -, — X0 = v2 = 0, or dn
Ri + Ri
, _ h _
a 12 — — » *®i — — yx = 0, or d12 = 0
Therefore
-1 2« t r 2RiR2t —i r
R,2
H(0 = \j <*i + -R2)C
i i
+ (Ri + R2)L
i -l
(Ri + R*)2C (*i + *2)2L
u0(t) T O'
+
Ri + R2 0 1
Simulation Difficulties
y&i y<i
CO
Y = (5.11-7)
_VvV\ VvV* • • • Vv .
384 State Variables and Linear Continuous Systems
v(t)
► Hift r) ^y(t)
t= 0
(a)
(b)
Fig. 5.11-1
Equation 5.11-8 can be used to evaluate the mean square outputs of the
system, assuming that H^r, t) and the autocorrelation matrix vCr^v2^)
are known, and that the integrations can readily be performed.
In many cases, Eq. 5.11-8 is too complex to permit analytical evaluation.
In the case of time-varying systems for example, it is frequently impossible
to determine a closed-form expression for Hr). In such cases, Y can be
determined by simulation techniques which have the advantage of not
requiring random signal generators.21 These simulation methods are
simplest when Y is to be determined at a fixed instant of time. This is the
case considered in the remainder of this discussion.
The basis for making random signal generators unnecessary in the
simulation is that Eq. 5.11-8 depends upon the autocorrelation matrix of
the random input, rather than upon the random input itself. Thus Y is
the same for Figs. 5.11-1# and 5.11-16, if the shaping filter is chosen so
that the autocorrelation matrix of v is the same in both cases. With
respect to Fig. 5.11-16,
(5.11-10)
Fig. 5.11-2a
Sec. 5.11 Impulse Response Matrices 387
Fig. 5.11-2b
h(t, r)
<xr(T)x(T) = (X:z’(t)B(t)v(t) dr
J — 00
where the cis’s are the indicated elements of the system C matrix. Then
t The term system is used in this section to denote the combination of the shaping filter,
switch, and system of Fig. 5.11-16.
t The D matrix is assumed to be a null matrix, since this is true for any practical com¬
binations of shaping filters and systems. Otherwise an infinite y\T) would result.
Sec. 5.12 Modified Adjoint Systems 389
ca(T)
cdT)
a(T) = (5.12-2)
CirtT)
oo
Y = H(T, T - rx)N(T - t,)Ht(T, T - Tl) dr. (5.12-4)
cn(T)
cdT)
“(0) = (5.12-6)
_CiniT)_
(a)
(b)
Fig. 5.12-1
cn(T - h)
ci2(r ^l)
JiJJ - h)_
Since B and C are, respectively, the input and output matrices of the
original system, the net result of this argument is that the ith row of
H(7", T — t) required for Eq. 5.12-4 is generated for variable r by reversing
the inputs and outputs of each of the simulation elements of
x = A (t)x + B (t)\
y = c(t)x
and replacing the time t of any time-varying gains by T — tv The resulting
system is called the modified adjoint system. It is compared with the
original system in Fig. 5.12-1.
Sec. 5.12 Modified Adjoint Systems 391
If the original system has m inputs and p outputs, the modified adjoint
system has p inputs and m outputs. Each simulation run on the modified
adjoint system produces m outputs. These m outputs comprise a row of
H(r, T — r), where the running variable of simulation t is equal to r.
The particular row of the matrix is determined by the input on which the
unit impulse is placed at the beginning of the simulation run.
Example 5.12-1. For the differential equation y + ty + y = 0, the curves of Figs.
5.11-2a and 5.11-26 show the cross-plotting operation required to obtain h(T, r). Show
that the modified adjoint system produces the desired result in one simulation run.
The impulse response of the system is given by
= 0 T > t
If T is substituted for t and the change in variable, r = T — t1} is made, then the im¬
pulse response h{T, T — rx) is
pT
h(T, T - rj) = e(A2-T2)/2 di Ti>0
JT-rx
= 0 TX < 0
The simulation diagram of the original equation is shown in Fig. 5.12-2a. Thus the
modified adjoint system has the diagram shown in Fig. 5.12-26. The differential equation
corresponding to Fig. 5.12-26 is
a + (T — ?i)a = 0
(a)
(b)
Fig. 5.12-2
392 State Variables and Linear Continuous Systems
Then
Let (3 = T — A. This yields /**(?!,()), the impulse response of the modified adjoint
system, as
rT
h*(h, 0) = a(0 = €U2-t2)/2 ^1, tx>0
jT-ty
This is identical with the expression for h(T, T — tx) if tx is written for tx.
Thus the impulse response of the modified adjoint system observed over the tx axis
is exactly the impulse response of the original system, observed over the rx axis. However,
the tx axis here is the time axis of observation on the computer, while the rx axis is the time
axis of application of impulses. Therefore a simulation run on the tx axis produces
exactly the same results as does a cross-plot sketched on the rx axis. This is shown in
Fig. 5.12-3, where a series of runs of h*(tu 0) are shown. Note that, if the r axis in Fig.
Sec. 5.12 Modified Adjoint Systems 393
5.11-2b were changed to r1 = T — r, these curves would be identical with the curves
of Fig. 5.12-3. Therefore the modified adjoint system produces in one simulation
run what would take many simulation runs of the original system.
" 0 1 “ eft) 0
, B = , c =
_bft)_ 0 c2(0_
The simulation diagram for this system is shown in Fig. 5.12-4a. Determine the diagram
for the modified adjoint system.
Since H(t, r) = C(t)<$>(t, t)B(t), the order of H(7, r) is (2 x 1), or two rows and one
column. The application of a unit impulse at time r = r2 produces an output on both
yx and y2, thus giving all the elements of H(7, t*), since H(f, r) is a single column matrix.
However, this is for variable t and fixed r = t*.
(a)
(b)
Fig. 5.12-4
394 State Variables and Linear Continuous Systems
If the inputs and outputs of the simulation diagram are reversed and the change in
variable t = T — tx is made, the result is Fig. 5.12-46. Note that at each terminal the
number of inputs and of outputs are interchanged. This fact provides a rapid partial
check of the correctness of an adjoint simulation diagram. The vector-matrix differential
equations for this system are given by
0 1 cifT 11) 0
[ax a2] = [ax a2] + hr r2]
a2\(T tf) a22{T tf) 0 c2(T - tf)
MT
q = [ai a2]
bfT
or
aT = ctTA(T - tf) + rTC{T - tf)
q = *Tb(T - L)
If a unit impulse is placed on one of the inputs at time zero, the output q{tf represents
a row (one element in this case) of the original response matrix H(T, T — tf). Therefore
two simulation runs are required to produce both rows of the impulse response matrix.
REFERENCES
15. B. Friedman, Principles and Techniques of Applied Mathematics, John Wiley and
Sons, New York, 1956, p. 43.
16. Coddington and Levinson, op. cit., p. 65.
17. Stuble, op. cit., p. 64.
18. Coddington and Levinson, op. cit., p. 80.
19. L. A. Pipes, J. Appl. Phys., Vol. 25, pp. 1179-1185.
20. M. Issac, “Deterministic and Stochastic Response of Linear Time-Variable Systems,
General Electric, LMED Technical Report, No. R62 EML1, March 1962.
21. J. H. Laning, and R. H. Battin, Random Processes in Automatic Control, McGraw-
Hill Book Co., New York, 1956, Chapter 6.
22. R. Sussman, “A Method of Solving Linear Problems by Using the Adjoint System,”
Internal Tech. Memor. No. M-2, Electronics Research Laboratory, University of
California, Berkeley, California.
Problems
(a) y + 3y + 5y = v
(b) y + 3y + 5y = v + 2v + v
5.2 What are the differential equations for the system shown in Fig. P5.2?
v\
5.3 Draw the simulation diagram for the following differential equations.
(a) y + ky + (co2 + e cos t)y = v (Generate cos t by simulation.)
(b) y + y2y + 5y = v
5.4 Show that the simulation in Fig. P5.4 satisfies Eq. 5.5-3.
Fig. P5.4
Problems 397
5.5 Find the transfer function matrix and draw the transfer function diagram
for the systems described below. Comment on the number of integrators
required.
5.6 Write the vector matrix equations for the system shown in Fig. P5.4 in
terms of the state variables indicated.
5.7 Write the vector matrix equations for the systems of Problem 5.5.
5.8 Find the vector matrix equations for the following systems using the partial
fraction technique. Show that these equations can be determined from the
equations obtained by simulation techniques by a coordinate transforma¬
tion.
(a) y + 3y + 2y = v
(b) ‘y + 4y + 5y + 2y = v
(c) 'y + Ay + 6y + My = v
(d) y1 - IO2/2 + Vi = vx
Vz + 6y2 = v2
5.9 Find the general expression for the elements of the A(/) matrix for an
RLC network if the indicators and capacitors are functions of time.
5.10 Given the system defined by the time-varying differential equation
n— 1 n
Pny +
k=0
2 <*n-lit)pky =
*=0
2 Pn-k(t)pkV
-±1~ - 0 10 0 - xx "bp
±2 0 0 1 0 Xz bz
— +
• • •
. 1
A. an an—1 • • • “al- .xn_ A_
y = X! + b0v
398 State Variables and Linear Continuous Systems
where
b0(o = m
2 ^ r ln -f- in — i\
bi(t) = Pi(t) ~I I _ . ai_r_m(t)pmbr(t)
r=0 m—0 \ ^ l 1
5.11 Find the simulation diagram for the system defined by the time-varying
differential equation
(a) y + 4y + 3y = 0
(b) y + 2y + y = 0
(c) y + 2y + 2y = 0
(d) y + 4y + 6y + 4y = 0
5.15 Assume that a unit step is applied to the systems described by Problem
5.14u, b, and c. What is the output y(t)l
5.16 Assume that an input, cos cot, is applied to the system of Problem 5.14c.
What is the output y(t)l For what value of oj does y(t) have its largest
peak amplitude? Does y(t) ever become unbounded?
5.17 In the vicinity of the operating point (TV = 1, C = 1, 7? = 0) of a nuclear
reactor, it can be shown that the linearized dynamic equations are given by:
x1 = — x± + x2 + x3
ry* - ry* _ ryt
x3 = Kxx
If all the A/s have negative real parts, show that in the steady state the
peak amplitude of oscillation of the zth mode is given by
<Xi, g)
jW - A;
2 eMut>(u
i=1
(b) Since x(7) = eAt x(0), for an unforced system, show that x(/)is equal to
2 x(o))cA<<uf
i=l
5.26 (a) Show that the state transition matrix eM can be found by the use of
the modal matrix as ext = where A is a constant matrix. Why
is the transformation x = M(/)q not particularly useful for the system
x = A (t)x?
(b) Using the result of part (a), show that the complete solution for the
matrix differential equation x = Ax + Bv is given by
x(0 = I
J— oo
MeA(^-^M-1Bv(A) dX
(c) Using the result of part (b) and the fact that the columns of M are the
characteristic vectors and the rows of M_1 are the reciprocal basis
vectors rif show that
rt n
x(0 = 2 (ri’ Bv(A))eA^-A>U; dX
J— 00 i=1
5.27 The transform of the state transition matrix of a fixed system is given by
= fsl — A]-1. Why is the inverse state transition matrix transform
<E>_1(s) not equal to [si — A] ?
5.28 For the system shown in Fig. P5.28, solve for the initial condition response
x(t) using the mode expansion technique, and check using the state transi¬
tion matrix.
Fig. P5.28
5.29 Consider the constant resistance network shown in Fig. P5.29, where
L(t)/C(t) = R2; L(t), C(t) > 0. Let v(t) be a voltage source, xL be the flux
linkages of the inductor (x1 — LiL), x2 be the charge across the capacitor
(x2 = Cvc), y(t) be the current to the network. Choose as state variables
xi = (^i + #2)/2» x2' = (aq — x2)/2. Write the differential equations for
x{ and x2' and show that x/ is Sc and x2 is S°.
Problems 401
5.30 Consider the following two subsystems, both of which are observable and
controllable:^ = ol1x1 + v(t), x2 = cc2x2 + v(t). If these two systems are
connected in parallel, such that the output y = x1 — x2, what are the
conditions required for the overall system to be both observable and
controllable?
5.31 Given the cascade connection shown in Fig. P5.31, show that8
(a) n = na + nb
(Jd) A^, . . . , Aw . . . , Ana, Ajj,, . . . ,
(c) A necessary (but insufficient) condition for the controllability (observ¬
ability) of S is that both Sa and Sb be controllable (observable).
(d) If Sa and Sb are both controllable (observable), any uncontrollable
(unobservable) modes of S must originate in Sb(Sa).
Fig. P5.31
5.32 Given the parallel connection shown in Fig. P5.32, show that8
(a) n = na + nb
(.b) Aj, . . . , An Aja, . . . , Ana, Ajj,, . . . , Anb
Fig. P5.32
402 State Variables and Linear Continuous Systems
5.33 Show that the transfer function H(.s') of a system S can be written as8
n*
H(i) = 2 + D
i=1
where the * indicates S*, and the matrices have rank one. Hint: Start
with the expression H(s') = COB + D where O = [jl — A]-1.
5.34 Consider the system
'-3 r ‘1 r V "i r
x + v,' y =
1 — 3_ _1 i_ j -i_
x = Ax + Bv
y = Cx + Dv
C = [1 -2]
(a) Find the region of state space for which the system is controllable.
(b) Find the minimum time required to return a state in the controllable
region to the origin.
Problems 403
Fig. P5.38
5.39 Find the state transition matrix for the following systems.
(a) y + ty — v
(b) y + y/t = V
(c) y + ty + y = v
(d) t2y + ty + y = v
x = A(t)x
then
5.46 For the equation y + p(t)y = 0, where p(t) is continuous with period T,
what are the conditions that this equation admit of periodic solutions?
5.47 A state transition matrix of a system satisfies the equation
5.48 Find the impulse response matrix for the systems of Problem 5.5.
5.49 Find the impulse response for the systems of Problem 5.14.
5.50 For the system with the following A, B, C, and D matrices, find the impulse
response matrix by use of the simulation diagram. Check your solution
using Eq. 5.11-5.
5.51 Find the impulse response matrix for the RLC circuit shown in Fig. P5.51.
5.52 Find the impulse response matrix H(7) for the system
1 1 0
y = 0 1 1
0 0 1
0 1 o ' '-l'
x = 5 0 2 x + 1
_ —2 0 —2_ -1
y = [—2 1 0] x
(b) Show that the intersection of the controllable and observable spaces
is the space generated by the transfer function matrix.
406 State Variables and Linear Continuous Systems
x = A (t)x + B(t)v
y = C(t)x
is given by
H(T, r) = c(D4>(r, t)B(t) t > r
= 0 T < r
((a) Show that the impulse response matrix of the adjoint system
= [0] TJ < 0
(c) From the result of part b, show that observation of the response of
the modified adjoint system over the tx axis (the running variable of the
simulator) is identical with observing the cross-plot at time T of the response
of the original system over the axis. In addition, show that an impulse
placed on one of the inputs to the adjoint system produces a row of the
desired impulse response matrix.
(d) Show that the result of the preceding parts proves that the complete
modified adjoint system is obtained by interchanging the inputs and out¬
puts of the original system and making the change in variable t = T — tv
6
State Variables and Linear
Discrete Systems
6.1 INTRODUCTION
407
408 State Variables and Linear Discrete Systems
Example 6.2-1. Find the simulation diagram for the system governed by the difference
equation y(JcT + IT) + ay{kT + T) + by{kT) = v(kT).
The first step is to solve for y{kT + IT), as
The terms y{kT + T) and y{kT) are obtained as shown in Fig. 6.2-2a. Assuming ideal
distortionless delays, a signal which appears at terminal 1 appears at terminal 2 one time
period later, and at terminal 3 two time periods later. Similarly, a signal at terminal 2
appears at terminal 3 one time period later. The completed block diagram (Fig. 6.2-2b)
is obtained by satisfying the requirements of the difference equation.
If the initial conditions are given in terms of y{0) and y{T), then y{0) is the initial signal
at the output of the first delay, and y(T) is the initial output of the second delay. After
one time period, y(T) appears at the output of the first delay unit. After two time periods,
the output of the first delay is y(2T) = n(0) — ay{T) — by{0).
If a comparison is made between Figs. 5.2-2b and 6.2-2'o, it is evident that similar
rules hold for constructing block diagrams of difference equations and differential
equations. The integrator used in simulating differential equations is analogous to the
unit delay used in simulating difference equations.
(a)
(b)
Fig. 6.2-2
Sec. 6.2 Simulation Diagrams 409
v(kT)
Example 6.2-2. Find the simulation diagram for the nth order difference equation
y(nT + kT) + a n-Xy{nT + kT — T) + • • • + oc0y(kT)
=finv(nT + kT) 4- fin-i v(nT + kT — T) + • • • + fi0v(kT)
The general simulation diagram for this system is shown in Fig. 6.2-3, by analogy to
Fig. 5.5-4. The a's and b’s of the block diagram are given by Eqs. 5.5-7 and 5.5-8.
For the specific case of the difference equation
y{kT + 3 T) + 3 y(kT + 2 T) + 4 y{kT + T) + y{kT) = 2 v{kT + 3 T)
+ 3 v(kT + 2 D + v(kT + T) + 2 v(kT)
b0 = fin = 2
b\ = fin-1 an-1^0 = ^
b% ~ fin-2 aw-l6i an-2^0 = 2
The simulation diagram for this system is shown in Fig. 6.2-4. A comparison of this
v(kT)
Fig. 6.2-4
410 State Variables and Linear Discrete Systems
diagram with Fig. 5.5-5 shows that the only difference between the two diagrams is that
the integrators of Fig. 5.5-5 have been replaced by unit delays.
Example 6.2-3. Find the simulation diagram for the system governed by the difference
equations
Using y2(kT + 2T) — v2{kT + T) as the input to one delay chain, the block diagram
appears as shown in Fig. 6.2-5. The approach is similar to that used for continuous
systems.
The transfer function H(z) for a single input-single output discrete time
system is equal to the ratio of the Z transforms of the output and input of
the system. For multivariable systems, the transfer function between
various input-output terminals is similarly defined. Thus
His(z)
m. (6.3-1)
J4(X) — ^5 k 7^ j
V,(z) ’
Sec. 6.3 Transfer Function Matrices 411
where Yfz) is the Z transform of the output at terminal i, and Vfz) is the Z
transform of the input at terminal j.
The transfer function matrix is simply the ordered array of these
transfer functions, where i denotes the row and j denotes the column in
which Hij(z) appears. If the transfer function matrix H(z) is known, then
the output vector transform Y(z) is given by
where V(z) is the column matrix of the Z transform of the input vector
v(kT).
Example 6.3-1. Find the transfer function matrix for the two input-two output system
described by the difference equations
Taking the Z transform of both sides of these equations, assuming zero initial con¬
ditions,
o3 + 6z2 + 11a + 6)^(2) = (z + 1)^(2) + v2(z)
(Z2 + 5 z + 6) Y2(z) =(z + 1 )V2(z)
1 1
H 11(f) = z \ zrr, : 4: Hl2(z) = -
(2 4- 2)(2 + 3) 12V ; (2 4- 1)0 + 2)(2 + 3)
2 + 1
H2l(z) = 0 H22(z) = -
22V ; (2 + 2)(2 + 3)
or
1 1
(2 4- 2)(z 4- 3) (2 4- 1)(Z + 2)(2 + 3)
H(2) =
(g + 1)
0
(2 4- 2)(2 + 3)
VT (z)
Fig. 6.3-1
412 State Variables and Linear Discrete Systems
Fig. 6.3-2
Example 6.3-2. Find the transfer function matrix for the system governed by the
difference equations
yx{kT + IT) + yx{kT + T) + y2{kT + T) = vx(kT + T) + vx{kT) + v2{kT)
y2{kT + T) 4- yx(kT) = v2(kT)
Transforming both sides of these equations, assuming zero initial conditions,
(z2 + z)Y1(z) + zY2(z) = (z + 1)^(2) + v2(z)
Yx(z) + zY2(z) = V2(z)
Solving for Yx(z) and Y2(z),
(2 + 1) V2(Z) (2+1)
Y1(z) = —-V1(z) T2(2) = Vi(z)
Z2 + 2 — 1 z(z2 +2—1)
_2(22 +2—1) 2_
Fig. 6.3-3
Sec. 6.4 The Concept of State 413
v = negative
input
pulses,
amplitude
Vo
Fig. 6.4-1
414 State Variables and Linear Discrete Systems
Fig. 6.4-2
Boolean Identities
And Or
0-0 = 0 0 + 0 = 0
0-1 = 0 0+1 = 1
1-0 = 0 1+0=1
1-1 = 1 1 + 1 = 1
Complement
0= 1 1 = 0
Unit
AB
A(k+l) delay AW And
Or
AB +AB
Logical
inverter And
AB
Logical
inverter
Unit
g(k + i) delay
B(k)
Fig. 6.4-3
Sec. 6.5 Matrix Representation of Linear State Equations 415
Solving this equation in the standard fashion indicated in Section 2.12 yields
,k~\
C,
x{k) =
(51 -(- C‘>,
Therefore the capacitor voltage increases in discrete steps as shown in Fig. 6.4-2. The
step heights are given by
Cx C2
x(k + 1) — x{k) = V0
A -f- CJ yCx + C2
The state of this circuit is given by the capacitor voltage x{k). Note that this state
takes on only discrete values. The next state of the system, x(k + 1), is uniquely deter¬
mined by the present state of the system and the present input. The general state
equations for this system are given by Eqs. 6.4-1. In this case, both f and g are linear,
single-valued, scalar functions.
Example 6.4-2.b A modulo 4 counter is designed so that it cycles through the counts
00, 01, 10, 11, 00, . . . . The circuit for this device is shown in Fig. 6.4-3, and the
reader familiar with sequential circuits can verify that this circuit does indeed cycle
through these counts. The outputs of the delay elements represent the state of the
system, as well as the outputs of the system. Write the state equations.
The logical equations for this counter are normally written in the form
B«+1) = (BfT
where the superscripts denote time instants and are not exponents. This is an auton¬
omous sequential circuit, and the general state equations for such a circuit are
The general form of the state equations for a multivariable discrete time
system were given by Eq. 6.4-1 as
If the system is linear, then Eq. 6.5-1 can be written as the set of linear
vector-matrix difference equations
xi(kT)
y(kT) = [ 1 0]
x2(kT)
The variables a;1(/:7)and x2{kT) are the outputs of the delay elements, and they represent
the state of the system. These equations are of the general form of Eq. 6.5-2, where
0 r "0“
A = , B = [1 0], D = [0]
—6 a 1
Example 6.5-2. The general form for an nth order difference equation is given in
Example 6.2-2. The simulation diagram appears in Fig. 6.2-3. Find the A, B, C, and
D matrices.
Analogous to the continuous case, the A, B, C, D matrices for this system are given
by Eq. 5.5-9.
Sec. 6.5 Matrix Representation of Linear State Equations 417
xfk + 1) = x2(k)
x2(k + 1) = xz(k)
xz(k + 1) = Xi(k)
xjjc + 1) = x^k) © xz(k) © xx(k)
In matrix form,
"0 10 0 "
0 0 10
x(k + 1) = x(/c) = Ax(k) mod 2
0 0 0 1
10 11
The sequence of states through which this network will pass is x(0),
Ax(0), A2x(0), .... If A is nonsingular, then each state x(A:) has a unique
preceding state x(k — 1). For a mod 2 network, the determinant |A| is
either one or zero. In this particular example |A| = 1, so that A is
Fig. 6.5-2
f Readers completely unfamiliar with sequential networks should omit this section.
418 State Variables and Linear Discrete Systems
0 0 1 1 0 1 0
—> —>
0 1 1 0 1 0 0
i 1 0 1 0 0 0
Sec. 6.5 Matrix Representation of Linear State Equations 419
Fig. 6.5-3
0 10 0
0 0 10
x(k + 1) = x(k)
0 0 0 1
110 0
|AI — A| mod 2
= A4 4- A3 + 1 = 0 mod 2
The partial fractions technique for deriving the state equations for a
linear, fixed, discrete time process follows the same procedure as that
used for continuous systems. For a single input-output system with
420 State Variables and Linear Discrete Systems
transfer function H(z), the transform of the output Y(z) is given by Y(z) =
H(z)V(z). If the denominator polynomial of the transfer function H(z)
has distinct roots Xl9 A2, . . . , An, and if the order of the numerator poly¬
nomial of H(z) is less than the order of its denominator polynomial, then
H(z) can be written as the partial fraction expansion
n c-
m = 2 -A-
i=i 2 — At
The output Y(z) is given by
n*) = 2i —^
z — Ai
v(b
Therefore y(kT) can be written as a sum of terms of the form
n
y(kT) = 2 cixi(kT)
2=1
where the xjiJcT) must satisfy the first order difference equation
• 0 0 • • • x
,'"n
*„[(* + 1)71 _xn(kT)
xi(kT) (6.5-3)
x2(kT)
[y(kT)] = [Cl cs cJ
xn(k T)
or
x[(fc + 1)J] = Ax(kT) + B v(kT)
(6.5-4)
y(kT) = Cx(k T)
make concise statements about the properties of the system. This form is
particularly convenient when dealing with forcing functions, as the corre¬
sponding state transition matrix is also of diagonal form. However, there
does arise a computational difficulty, because the transfer function H(z)
gives no information about the initial conditions of the system. In fact,
to compute the initial conditions on the x/s in this form, one must find
y(0), y(l), . . ., y(n — 1) and then solve a set of simultaneous equations to
find the relationships between the boundary conditions on the yf s and those
on the state variables of Eq. 6.5-4.
Example 6.5-3. Find the A, B, and C matrices for the sampled-data system of Fig.
6.5-4 a.
With respect to the sampled input and output, the transfer function H(z) of this
system is given by
Y(z) , 1
—— = H(z) = ---—=
V (z) 2 — 1 2 — e al
V(z) V(z)e~aT
Y(z) = H(z)V(z) = -!-4- ■ = Xt(z) + X2(z)e~°T
2 — 1 2 — e~al
(a)
O.
\f —
v(kT)
r~aT
Q -^y(kT)
J+
(b)
Fig. 6.5-4
422 State Variables and Linear Discrete Systems
H(z) = -5- = 1 +
2—1 1
This operation is equivalent to
D = D = lim H(z)
Z~* 00
where A is the diagonal matrix whose elements are the characteristic roots
A, k2, . . . , Xn of the denominator polynomial of H(z), and
B = column matrix whose elements are equal to one
C = row matrix all of whose elements are equal to ci = (* - K)m\z^
D = d{) — single-element matrix = lim H(z)
2—>00
Example 6.5-4. Find the A, B, C, and D matrices for the system whose transfer function
is
4z3 _ 12Z2 + l3z - 7
= —7-7^7-
(z - 1 )2(z - 2)
c2 = ^
dz
[(2 -1 mow = i
c3 = (z- 2)H(z)jz=2 = 3
Fig. 6.5-6
For the multivariable case where there are multiple inputs or outputs,
the transfer function matrix H(z) can be used in a similar manner to find
the state equations. However, the approach is not so clear-cut as in the
single input-single output case, as there is generally a greater freedom of
choice in assigning elements to the B and C matrices.
Example 6.5-5. Find the A, B, C, and D matrices for the system whose transfer function
diagram is given in Fig. 6.5-7.
The outputs Fi(z) and F20) are given by
z 1
Y1(z)
0 + 1 )(z + 2)
FiO) + V2(z)
2 + 2
1 2
Yi(z) = - K(z) + Vx(z) + V2{z)
z + 1 Z + 2 2 + 2
9 .
i 2
Y2(z) = V1(z) ~ V2(z) + VM + 4 no)
z + 1 2 + 3
Sec. 6.5 Matrix Representation of Linear State Equations 425
V2(z)
Using these equations, Fig. 6.5-8 was drawn. It is readily apparent that the corre¬
sponding state equations are
Fig. 6.5-8
426 State Variables and Linear Discrete Systems
where Bn = M-1B, the normal form input matrix, and Cn = CM, the
normal form output matrix. This is a general procedure, and, since it
originates from the difference equations of the system, it is to be preferred
over the transfer function matrix approach.
Example 6.5-6. The simulation diagram of Fig. 6.5-9 represents the original system
whose transfer function matrix was given in Example 6.5-5. Determine the state
equations in normal form.
Sec. 6.5 Matrix Representation of Linear State Equations 427
vi(kT) V2(kT)
~xx(k T)~ r-
~yi(kT)' 1 O' ~vx(kT~
xi(kT)
_y*(kT)_ 0 ~ 4 _v2(kT
MkT)_
For the A matrix above, the modal matrix M, and its inverse M1 are given by
Substituting into Eq. 6.5-8, the resulting normal form matrices are
A check of all the transfer products bijCki shows that this set of matrices represents the
same system as that of Example 6.5-5.
The new state variables q are related to the old state variables x by the relationship
q = M-1x, or
qi = hxi q* = —(*1 + x*) q3 = ixi + *3
428 State Variables and Linear Discrete Systems
The use of the relationship q = M_1x removes the previous difficulty of finding the
initial conditions on the state variables q in terms of the known system initial conditions.
Finding the modal matrix may involve no more labor than any of the
other methods for finding the state transition matrix. In view of the
advantages of having a diagonal A matrix, the normal form is quite
desirable. It is also interesting to note that the mode expansion technique,
to be considered next, and the normal form produce the same effect of
uncoupling the state equations. In this respect, they are identical
approaches but are written in different forms. The form in which the
system equations are expressed is frequently one of personal preference
and familiarity.
The concept of expanding the response of a linear fixed system into the
sum of responses along the characteristic vectors of the A matrix can also
be applied to discrete systems with distinct characteristic values. The
development follows directly from the equations in normal form.
From Eq. 6.5-8,
q (T) = Aq(0) + B„v(0) (6.6-1)
and
q(2 T) = Aq(D + B XT) (6.6-2)
Substitution of Eq. 6.6-1 into Eq. 6.6-2 yields
M = [uj u2 • • • u„]
Sec. 6.6 Mode Interpretation 429
n _
x(kT) = [ux u2 • • •
u„]A*
**n x(0)
T
■n
k-1
■ n
This can be rewritten as
Fig. 6.6-1
A =
The characteristic roots are = —2, and A2 = —1. The normalized characteristic
vectors are
1/V2 I/V5
iii = u> =
-1/V2 -2/V5
The reciprocal basis is then
1_
1
1m
>
2V2
1
12 —
1_
<1
_-y/5_
to
1
The normalized modal matrix M and its inverse Mr1 are given by
1
>
Note that the rows of M-1 are the reciprocal basis vectors r,.
The scalar products (ri5 x(0)> are
<rx, x(0)> = 2V2 + V2x2(0)
<r2,x(0)) = — V5x!(0) - V5 «2(0)
Sec. 6.7 Controllability and Observability 431
2 V2 ^(0) + V2 z2(0)
q(0) = M-1x(0) =
-V5^(0) - V5x2(0)
The forcing function Bv(kT) is the vector b times the scalar v(kT), where
b =
fc-i
x(kT) = [2V2*1(0) + V2z2(0)](-l)*u1 + 2 V2 (-1 )<^-1)y(yT)u1
3=0
k—1
+ [-V5^(0) - Vs z2(0)](-2)*u2 + Ji(-V5)(-2)t'-,-1'v(jT)tt2
3=0
The general expression for the time response q(kT) is, from the above and Eq. 6.6-4,
k-1
(-1)*?1(0) + 2 V2 (-l)IM-I»t)(;T)
i=o
q (kT) =
A:—1
(-2)*?!(0) + 2(-V5)(-2)<*-)-1,b(;T)
3=0
then all the modes are controllable if there are no zero rows of Bn. Stated
in terms of the mode expansion method, this means that none of the scalar
products <rf, Bv) vanishes.
Observability is a function of the coupling between the modes of the
system and the output of the system. All the modes of the system are
observable if there are no zero columns of Cn.\ Alternatively, this require¬
ment could be stated as: The kih mode is not observable if all the scalar
products (c*, uk) = 0 for all Vs, where the vector c* constitutes the zth
row of the original C matrix.
For a sampled-data system there is an additional requirement. If the
continuous system has a partial fraction expansion which contains the
term p/[(s + a)2 + p2], and if the sampling interval T = ir/p, then the
Z transform of this term,
og _r_
= 0
2-2 aT
L(s + of + p\ 1 - lz~\-aT COS PT + Z-V
The system may even be unstable, with a < 0, but this fact could not be
inferred from observations of the output. These are called “hidden oscilla¬
tions,” and they occur when the zeros of the oscillation coincide exactly
with the time that the system is sampled.10-12 In this situation the system
is neither completely controllable nor completely observable. Therefore
the additional requirement for complete controllability and observability
of sampled-data systems is that, if a characteristic root of the continuous
system is — a ± jfi, then T ^ v/p.
The state transition matrix for the linear discrete time system is inves¬
tigated in this section. Similarly to the continuous case, the state transition
matrix is the fundamental matrix of Eq. 6.8-1 below, subject to the condi¬
tion that<|>[(/c0, k0)T] = I, the unit matrix. Consider, then, the time-vary¬
ing state difference equation
Similarly,
x[(k0 + 2)T] = A l(k0 + l)T]A(k0T)x(k0T)
= I (k = fe0)
This is analogous to the continuous case, where the solution for a fixed
system depends only upon time differences; whereas for a time-variable
case the solution depends upon both the time of application of cause and
the time of observation of effect.
For the time-varying case where A (kT) can be written as the sum of two
matrices A0 and Ai(kT), a perturbation technique can be used to obtain
the state transition matrix. This procedure is useful if the time-varying
matrix Ax(kT) represents a small perturbation upon the constant matrix
A0. For this case
x[(k + 1 )T] = [A0 + A ^kTftxQcT) (6.8-6)
The solution x(kT) for the system of Eq. 6.8-7 can be determined by the
same process of iteration used to give Eqs. 6.6-4 and 6.8-2. Thus
The state transition matrix for discrete systems has a set of properties
which are directly analogous to the properties listed for a continuous
system in Section 5.9. Namely,
4>[(k0, k0)T] = I (6.8-12)
4>[(*2, ko)T] = <1>[(*2, K)T) (6.8-13)
4>[(*i, K)T] = 4>-i[(fr2, kx)T] (6.8-14)
For fixed systems,
<\>[(k + n)] = <j>(A-)c|>(«) (6.8-15)
<|>(A:) =4>~1(-^) (6.8-16)
Sec. 6.8 The State Transition Matrix 435
Computation of c|>
In general, the computation of the state transition matrix for the time-
varying case is a formidable task. Clearly, for any large value of n,
Eq. 6.8-4 becomes most unwieldy. In certain cases, where the difference
equations of the system can be handled, an analytical solution can be
obtained (see Section 2.13). However, this occurs rarely. Use of a com¬
puter is generally the best method to obtain a solution.
For the fixed system, an analytical solution can generally be obtained.
Equation 4.10-20 provides one method of computing the state transition
matrix. Some others follow.
A =
assuming x-dJcT) = y(kT), x2(kT) = y[(k + 1)7]. The characteristic equation |AI — A|
= 0 has two characteristic roots Ax = —2, A2 = —3. Therefore
F(AX) = Ay = (—2)k = a0 -f oqAj = a0 — 2cc1
F(A2) = Ay = ( — 3)k = a 0 + ax/l2 = a0 — 3ax
From these two equations, a0 = 3(—2)k — 2(—3)k and oc1 = (—2)k — (—3)k. Hence
F(A) = Ak = a0I + axA
or
' 3(—2)fc - 2(—3)* (-2Y-C-3)k -
<KU =
_—6[(—2)fc - (-3)*] —2(—2)k + 3( — 3)fc_
Example 6.8-2. Compute <t>(£) for the system whose A matrix is given by
A =
where is the order of the root. Hence ( — l)fc = a0 + cclA1 — a0 — ax and —k(—l)k —
ax, or ax = — k{— l)fc and a0 = ( — l)fc(l — k). Therefore
ao al —k ~
<t>(A) = a0I -(- axA =
-«1 a0 — 2ax 1 + k
436 State Variables and Linear Discrete Systems
Example 6.8-3. It is informative to take up the case in which the A matrix may have
complex roots. For this reason, determine <f>(&) corresponding to
0 r
A =
-2 2
The characteristic values of this matrix are 1 ± j. For purposes of evaluating Afc, the
polar form V2 e±^/4 is most useful. The computation of Afc is then
Adding and subtracting these equations yields (2)fc/2 cos (Jar14) = a0 + ax and ax =
(2)fc/2 sin (/C77-/4), or a0 = (2)fc/2[cos (kir/4) — sin (/T77-/4)]. Since
a0 ai
<4>(A-) = a„I + axA =
— 2ax a0 + 2ax
then
<J>(T) = (2)fc/2
(kir kir
cos-b sin —
4 4
Transforming both sides of Eq. 6.8-17, zX(z) — zx(0) = AX(z), where use
has been made of Eq. 3.12-3. Thus
This form is slightly different from the analogous form ^(s) = (si — A)-1
for continuous systems. The state transition matrix is given by the inverse
z transform of 4>(z), or
Example 6.8-4. Using the A matrix of Example 6.8-1, determine <|>(T) by the fre¬
quency-domain method.
Sec. 6.8 The State Transition Matrix 437
Example 6.8-5. Using the A matrix of Example 6.8-2, determine <\>(k) by the frequency-
domain method.
For this case,
"z -1 “
(zl - A)
1 z -f- 2
so that
"z + 2 r
-l z_
(zl - A)-1
(Z + l)2
—k
c|>(/c) = (~l)k
1 + k
If all the state variables except the yth are set equal to zero, and a unit
initial condition is placed on xjt then the response at the ith state variable
xfkT) represents the term ^ifk). Therefore the transfer function from the
output of the yth delay to the output of the zth delay represents the term
O^(z). This is slightly different from the continuous case, where the
transfer function was calculated between the input to they'th integrator and
the output of the zth integrator. The difference is due to the fact that the
transfer function in the continuous case is expressed as the transform of the
438 State Variables and Linear Discrete Systems
impulse response (in the continuous case, a unit impulse input establishes a
unit initial condition immediately), while in the discrete case it is the trans¬
form of the unit initial condition response. If an analogy is desired, then
the transfer function from the input of the yth delay to the output of the zth
delay must be multiplied by 2 to obtain Oi5(z). Since what is generally
desired is (zl — A)-1 = z-1<i>(z), the transfer function from the input of
the yth delay to the output of the zth delay is perfectly suitable.
Example 6.8-6. The simulation diagram of Fig. 6.8-la represents the system of Example
6.8-1. Determine <&(k).
Figure 6.8-16 is the same diagram redrawn for the convenience of computing transfer
functions. Since 1/(1—loop transfer function) is equal to z(z + 5)/(z2 -F 5z -f 6), the
(a) (b)
Fig. 6.8-1
various transfer functions z~1<&ij(z) can be obtained by multiplying the forward transfer
function from j to i by z(z + 5)/(z2 + 5z + 6). By performing this operation, the matrix
(zl — A)-1 = z_1<i>(z) is obtained. Thus
z + 5 1
—6 z_
[zl - A]-1 = z~'®(z)
z2 + 5z + 6
f In the case of complex elements, the transpose is replaced by the conjugate transpose.
Sec. 6.8 The State Transition Matrix 439
Note that the presentation here is the reverse of that of Chapter 5. There
the continuous analogs of Eqs. 6.8-28 and 6.8-30 were defined, and the
continuous analog of Eq. 6.8-23 resulted. Here Eqs. 6.8-28 and 6.8-30 are
obtained from the definition of the adjoint operator, Eq. 6.8-23.
For the discrete time adjoint system, the state transition matrix
{♦"'[(ft, K)T]}T
can be found by iteration of Eq. 6.8-28. The result is
The state transition matrix for the original system, for k0 < k, can be
found by reverse iteration of Eq. 6.8-1, i.e., Eq. 6.8-1 can be written as
x(nT) = A~\nT)x[{n + 1)7"], and this expression can then be iterated
from n = k down to n — k0. The result is
4>-1[(k,k0)T]=<S>[(k0,k)T] k> k0
-1 (nT)
4>Po, k)T =n=k
VA o
k0)T\ = n A (nT)
ko ♦ +k
k-1
[(ko, k)T] = n A(nT)
n — ko
^ k~l -1
<)>-1 \(k, k)T] = n A l(nT)
n=kn
Fig. 6.8-2
Sec. 6.9 The Complete Solution 441
Note that the form of the state transition matrix for increasing time
(Eq. 6.8-4) is different from the form for the state transition matrix for
decreasing time (Eq. 6.8-32). The reason for this difference is that reversing
the time direction for a discrete system entails an inverse A matrix
{x(kT) = Ar\kT)x[{k -f 1)T]}. Reversing the time direction for a con¬
tinuous system simply entails reversing the sign of the A matrix [<d/d(—t) =
—dldt]. A set of alternative forms of the state transition matrix is shown
in Table 6.8-1.
Table 6.8-1
If k is replaced by m in Eq. 6.9-3 and both sides are then summed from k0 to
k — 1, the result is
The first term on the right side of Eq. 6.9-5 represents the initial condition
response of the system, while the second term represents a superposition
summation of the effects of the forcing function. This equation is analo¬
gous to Eq. 5.10-6 for the continuous case.
When the system under investigation is fixed, Eq. 6.9-5 can be written as
the sum of an initial condition response and a convolution summation:
k—1
(6.9-8)
for a fixed system.
For the case in which the system is fixed, it is frequently convenient to
use the frequency-domain approach. For this case, Y(z) can be found by
transforming Eq. 6.9-1 directly. The result of this operation is
X(z) = (zl - A)-1zx(0) + (zl - A)-1BY(z) (6.9-9)
An example is now given which illustrates the various approaches that can
be taken.
1. Classical Solution—Time Domain. Assume yH(k) = ftk. Then (/?2 + 5(3 + 6)j3k =
0, or /h = —2 and /?2 = —3. The homogeneous solution is then yH(k) = Ci(—2)k +
C2( — 3)fc. Assume that yP(k) = C3. Then C3 + 5C3 + 6C3 = 1, or C3 = A. The
total solution is
y(k) = Cx(—2)fc + C2(—3)fc + A
Substituting these constants into the total solution gives the complete solution
B = C = [1 0], D = [0]
1
The output y(k) is given by Eq. 6.9-8 as
k—1
y(k) = C<t>(A:)x(0) + 2 C4>[(^ -m - \)]By(mT) + Dy(kT)
m=0
which, for this case, reduces to
Jc-l
y(k) = <f>u(k)y(0) + ^(l) + J 0i2(k - m - 1)
m=0
The computation of the summation for a forced input may involve some skill in
finding a closed-form expression for the resulting series. The summation formula
ki 1 — ak
X ak-m-1 _ -- (sum of a geometric series)
m=0
1 — a
1 _ (—2)k 1 — ( —3)fc 1 1 1
-_1 - u - -—L_iL = - + - (-3)fc - - (-2)*
1 - (-2) 1 - (-3) 12 4 3
444 State Variables and Linear Discrete Systems
The expression above for the sum of a geometric series is particularly helpful if the
matrix summation
k-1
TO— 1
m=0
N N
^ u(/c) Av(k) — [uij^vik)]1^1 — ^ v(k +1) Au(k)
M M
k 1 pk — J pk
v(k) = ^ rn + Q = + Ci — + Co
r - 1 r-1
N 2T+1 1 N
±1
kC 1
^ krk =
o r - 1
1
- [NrN+2 - (A + 1 )rN+1 -hr] r ^ 1
(r - 1)
Ri = (z- \)Y(z)zk~1\z=1 =A
Rz = (z + 2)Y(z)zk~'\z=_2 = [-1 + 3t/(0) + y(m~2)k
+ [i(-3)s - K-2)‘+ A]
2 T 5 1
-6
(zl - A)-1 =
(2 + 2)(z + 3)
O'
B = C = [1 0], D = [0]
1
The transform of the output, Y(z), is then
x 2(2 + 5)
Y(z) = - — y(0) +
+ + 3)
y( i) +
(2 + 2)(2 + 3) (2 2)(2 (2 - 1)(2 + 2)(2 + 3)
This 2 transform corresponds to the Y(z) found in part 3 of this example. Therefore the
answer is identical with that given in the three preceding parts of this example.
2 + 5 1
Adj [21 - A] =
Therefore M is given by
' 1 r ' 3
M = and M 1 =
-2 -3 -2
Since
-2 O'
M XAM = A =
0 -3
q (k + 1) = Aq(*) + B nv(k) (T = 1)
y (k) = Cnq (k) + Dv(/:)
where Bn = M_1B and Cn = CM. Therefore qx = 3xx + x2, q2 — —2xx — x2, and
1
B,i = and Cn = [1 1]
-1
The output y(/c) is then
k-1
y (*) = C„A‘q(0) + J C„A‘-”*-1Bnv(m)
m=0
446 State Variables and Linear Discrete Systems
from Eq. 6.6-4 and the above. Since A is a diagonal matrix, finding Ak means simply
raising the elements of A to the k\h power. This is one of the advantages of using this
method. It follows that
or
m = (-2)^(0) + (-3)^2(0) + [K-3)fc - K~2f + A]
If it is desired that y(k) be expressed in terms of the original initial conditions, then
qi(0) = 3aq(0) + *2(0) = 3i/(0) + 2/(1) and qz(Q) = — 2aq(0) — s2(0) = — 22/(0) — 2/(1)
yield
y{k) = [32/(0) + 2/(l)](—2)fc - [22/(0) + 2/(l)]( —3)fc + [£(-3)* - K~2)k + A]
r, = and r2 =
1 -1
form a reciprocal basis. Thus, from Eq. 6.6-6,
This is the same expressions as for y(k) in terms of the </>’s in part 2. Thus performance
of the indicated summations and substitution of 2/(0) = z+0) and 2/(1) = «2(0) yields
the desired result.
z2 + 5z + 6 z + 2 z + 3
then
V(z) V(z)
Y(z) = —4 - 4 - = xx(z) + xt(z)
z+2 z+3
or y(k) = xx(k) + x2(k), where xx and x2 satisfy the first order difference equations
Since these are exactly the same matrices which were derived in part 5, y(k) is then
However, since the transfer function gives no clue to the relationship between y and
x, this relationship must be obtained from examination of y{k). This is necessary
because the known initial conditions are in terms of 2/(0) and y{ 1). The desired relation¬
ships can be found by substituting k = 0 and k = 1 into y(k) = xx{k) + x2(k). This
gives 2/(0) = 3^(0) + z2(0) and 2/(0 = aq(l) + #2(0* Now, using the equations for
x^k + 0 and x2(k + 0 at k = 0, these expressions become
which are the same relationships found in part 5 for the q's. Although this procedure is
not too difficult for a second order system, it may prove to be laborious for higher order
systems with many inputs and outputs.
For a discrete fixed system, the state transition matrix approaches zero
as k approaches infinity, if the characteristic values of A are located inside
the unit circle. If a characteristic value lies on the unit circle and is of
order one, then 4>(k) is bounded as k approaches infinity. For any charac¬
teristic values which lie outside the unit circle or for multiple characteristic
values which lie on the unit circle, <\>{k) becomes infinite as k approaches
infinity. These statements can be proved from the Cayley-FIamilton
method of obtaining A/c, which depends upon obtaining certain elements
oim such that
448 State Variables and Linear Discrete Systems
where n is the order of the A matrix. The am are obtained from the equa¬
tions
to—1
or
A comparison of Eqs. 6.10-1 and 6.10-2 shows that the unit function re¬
sponse matrix is
C(fcD4>[(k. k0 + l)T]B(/c0T) k ^ k0 + 1
H[(/c, k„)T] =
D (kT) k = k0
= [0] k < k0 (6.10-3)
k ^ k0 -f 1
k — ko
Sec. 6.10 The Unit Function Response Matrix 449
In the frequency domain, from Eq. 6.9-10, the unit function response
matrix H(z) for a fixed system is given by
r~ -y(kT)
v(t) 1 — e
— sT
1 ■
s(s + 1J
-y(t)
V
T= 1
Fig. 6.10-1
1
G(z) = (1 - z-1)^
s\s + 1) 2 — e”1 2 — 1
Using this transfer function, the sampled-data system of Fig. 6.10-1 can be redrawn as
the discrete time sy stem of Fig. 6.10-2. In this figure, the forward path has been broken
y(kT)
up into its partial fraction expansion. The closed-loop system state equations can then
be written down by inspection. They are
1_
Xx (k + 1) 1 - e—1_
m
rH
+
o
xx(k)
y(k) = [i i]
x2(k)_
450 State Variables and Linear Discrete Systems
Mk- i) <f>l2(k- ly V1 - 1
H(/c) = [1 1]
_4>2l(k 1) 4*22^ 1).
Since
a0 + ax ax — a^-1
<J>(A: — 1) = A(A:_1) = a0I + axA =
ai a0
then H(k) = a0e_1 + ax(l — e-1), where a0 and cn1 are to be determined. The roots of
the characteristic equation
|AI — A| = 22 — 2 + (1 — c-1) = 0
]\j(k—l)€j(k—l)0 = a0 + oi^ei0
]\f(Jc-l)e-j(k-l)6 = ao +
(1 e-i)N(*-2)
+ sin (& — 1)0
sin 0
or
#(£) = (O.795)(fc-O|o.368 cos [0.680(/c - 1)] + 0.724 sin [0.680(A - 1)]}
0.72 \
H(k) = 0 MSN'*-1' cos (k — 1)0 + ^cot 0 + sin (k — 1)0
N sin 0/
For a time-varying system, the elements of the unit function response can
be obtained by simulation in a manner analogous to that used for con¬
tinuous systems. For an m input-/? output system, the ith output y^kT) is
Sec. 6.10 The Unit Function Response Matrix 451
given by
k m
Theyth column of H[(/c, n)T] can be obtained by setting all the inputs except
theyth equal to zero, and placing a unit function on theyth input at time
nT. The outputs of the system are h^Kk, n)T] for i = 1,2,...,/? and for
fixed j. In order to obtain the complete response /zi;[(/c, n)T] as a function
of both k and n, a set of runs must be performed, each run starting at a
different time nxT, n2T,... . The results of these runs must then be cross-
plotted to obtain the variation with respect to nT, the point in time of
application. Proceeding to different values of /', these tests must be repeated
until all m columns of the unit function response matrix are obtained. This
is the same problem that was presented in the last chapter where the impulse
response matrix H(f, r) was obtained by a similar cross-plotting procedure.
In this discrete case, however, the difference equations can be solved on a
digital computer and the necessary cross-plotting also done by the com¬
puter. Thus the discrete modified adjoint system is not discussed.
Example 6.10-2. In Section 5.11, the differential equation y + ty + y = 0 was ana¬
lyzed, and the impulse response h(T, t) was obtained by performing a set of simulation
runs for different values of r and cross-plotting the results for fixed t = T. The results of
the simulation runs are shown in Fig. 5.11 -2a, and the results of the cross-plotting are
shown in Fig. 5.11-26. Perform the same task, but for the discrete version of the same
differential equation. This method is often used to solve a time-varying differential
equation numerically by either a desk calculator or a simple digital computer routine.
A discrete version of the equation can be obtained in the following manner. Since
y(t + h) — y{t)
h
V(k0T) = 0
y(k0T+ T)-y(k0T)
T
or y(k0T) = 0 and y(k0T + T) = T.
452 State Variables and Linear Discrete Systems
k T—>-
Fig. 6.10-3a
Using a value of T — 0.05 and 60 = 0, 4, 8, 12, , 48, the points shown on Fig.
6.10- 3tf are obtained. Obtaining these points is a relatively simple task for a digital
computer, and the cross-plotting can also be performed by the computer. A desk
calculator can also be used. However, care must be taken to use sufficient accuracy
in computing each point, as the round-off errors can build up rapidly. The interested
reader can consult any of the many texts written on numerical solution of differential
equations.14’15
As a comparison of the accuracy that can be obtained by simple numerical methods
the results of the continuous system simulation run and the results of the approximate
numerical solution for r = k0 = 0 are listed in Table 6.10-1. The numerical solution
is given to three places, while the simulation run is given to two places, this being the
accuracy of reading from the original recording. For this comparison, the discrete
simulation reproduces the results of the continuous simulation within the accuracy of the
recording of the continuous information.
A crossplot of the points at KT = 2.0 is shown in Fig. 6.10-36. The points of Fig.
6.10- 36 represent 6(2.0, k0T), the unit response of the system as kT = 2.0 as a function
of the time of application of the unit input. A comparison of the cross-plot obtained by
discrete simulation and the cross-plot obtained by the continuous simulation (Fig.
5.11- 26) is shown in Table 6.10-2. This comparison shows that the discrete simulation is
fairly good, but not within the accuracy of the continuous system. The differences
between the continuous and discrete systems are due to the accumulation of round-off
error and the basic approximation involved. Because these effects are more noticeable
farther out along a simulation run, a cross-plot at kT = KT shows these effects more than
a cross-plot at kT < KT.
Sec. 6.10 The Unit Function Response Matrix 453
Table 6.10-1
T = 0.05
0 0 0 22 0.751 0.75
2 0.100 0.10 24 0.766 0.76
4 0.198 0.20 26 0.771 0.77
6 0.291 0.29 28 0.769 0.77
8 0.380 0.38 30 0.760 0.76
10 0.461 0.46 32 0.746 0.75
12 0.534 0.53 34 0.727 0.73
14 0.597 0.60 36 0.705 0.71
16 0.651 0.65 38 0.679 0.68
18 0.695 0.69 40 0.653 0.65
20 0.728 0.72
Figure 6.10-36 can also be obtained by a discrete simulation of the modified adjoint
differential equation. The original differential equation is y + ty + y = 0, and the
adjoint differential equation is a — (d/dt)(t(x.) + a = 0 or a — to. = 0. Making the
change in variable t = T0 — tu the modified adjoint differential equation is then
d2o do.
+ (T0 — 11) = 0
~dtT2 dtx
0.7
1-T”
1
o
_ • • T = 0.05
0.6
•
•
0.5 _ •
•
P 0.4
Cross-plot of unit function •
o' response at KT = 2.0
C\j
^ 0.3 — • —
0.2 —
•
0.1 — —
_1_J__1 -1
0 0.5 1.0 1.5 2.0
kT-^
Fig. 6.10-36
454 State Variables and Linear Discrete Systems
Table 6.10-2
T = 0.05
0 0.653 0.65
4 0.628 0.63
8 0.603 0.60
12 0.574 0.56
16 0.541 0.53
20 0.501 0.48
24 0.450 0.43
28 0.385 0.37
32 0.297 0.28
36 0.175 0.17
40 0 0
The discrete simulation of this equation is found by the procedure used to find the
discrete simulation of the original differential equation. The resulting difference
equation is
where KT = T0 is the fixed end time. A comparison of the points obtained by this
simulation and those obtained from the continuous system simulation is shown in
Table 6.10-3.
Table 6.10-3
KT = 1.0, T = 0.1
k hjc'*(k', 0) 0)
0 0 0
1 0.100 —
2 0.191 0.18
3 0.276 —
4 0.350 0.33
5 0.420 —
6 0.487 0.47
7 0.550 —
8 0.612 0.60
9 0.673 —
10 0.733 0.72
Sec. 6.10 The Unit Function Response Matrix 455
Transmission Matrices
" h(0, 0) 0 0 0
h(T, 0) h{T, T) 0 0
Hx(kT, k0T) = h(2T, 0) h(2T, T) h(2T, 2T) 0
All elements to the right of the element h(iT, iT) are zero, since the system
is assumed to be physically realizable, or nonanticipative. If the input
v(kT) is ordered into a column vector whose components are i>(0), v{T),
v{2T),. . . , v(mT) then the output of the system can be written as
For a fixed system, the elements of the transmission matrix are h(iT —
jT), the argument being the difference between the time of observation and
the time of application. Thus
m 0 0 • • • 0
h(T) m 0 • • • 0
H T(kT) = h(2T) h{T) h{ 0) 0
Example 6.10-3. Repeat Example 1.9-1 using the transmission matrix description.
y(k), k = 0, 1, 2, ... , can be written as
y{k) = DT(k)\(k)
where
“2/(0)-
y{ i)
y(k) y{ 2)
and
^(0)
d{ 1)
d(/c) = d{ 2)
Example 6.10-4. Write an expression for y(kT), k — 0, 1, 2,... , for the system of
Example 3.13-6.
Defining the indicated transmission matrices and vectors as above,
+ VJ2^2 + * * * + Vjm^m = VJ
where the subscript on the y's and the first subscript on the v's denote a
particular set of measurements. The subscript on the /z’s and the second
subscript on the v's denote the argument (m + 1 )T. If only M sets of
measurements are made, i.e., J = M, unique solutions exist for the h^ but
noise and measurement errors generally cause Eq. 6.11-1 to yield incorrect
values. Thus, in such an experimental situation, many measurements are
usually taken, and more equations than the number of unknowns are found.
Hence J > M. Now, however, values for the ht cannot be found which
satisfy all the equations. For example, substitution of the values h-f,
h2°, . . . , hm° for the unknown hx, h2, . . . , hm in the left side of Eq. 6.11-1
might yield yf, y2°, . . . ,yj° which differ from yv y2, ... ,yj by et =
yf — y., i = 1,2,...,/. Faced with this dilemma, one might attempt
458 State Variables and Linear Discrete Systems
to determine the hi in such a way that each of Eqs. 6.11-1 is at least approxi¬
mately valid, and such that some measure of the total approximation error
is as small as possible. For example, the h*° can be determined such that
the ei have the smallest possible mean square deviation, i.e.,
/ = £*'
=1
i
= f (Vi9 - i—1
Vif
• •
Jj_
Thus the problem is to determine
hf
Sec. 6.11 The Method of Least Squares 459
Fig. 6.11-1
such that y° has the smallest possible deviation in norm from y. This is
represented for the case M = 2 in Fig. 6.11-1. In the general case, the set
of all linear combinations of y1} v2, . . . , vM forms a space R°, and the
orthogonal projection of y on R° is the vector in R° which is the closest to y.
This is a simple generalization of what is geometrically apparent for
M — 2 in Fig. 6.11 -1. Thus h° is to be chosen so that the linear combination
yo _ oYi hfv2 + • * * + hM°\M is the orthogonal projection of y on
R°.
Given a space R° and a vector y, which is in general not contained in
R°, y can always be represented in the form
y = y° — e
or
/q°(vi, Vi) + h2°(y2, yx) H-+ yx) = (y, vi>
v2) + h2°(y2, v2) + • • • + hM°(\M, v2) = <y,
(6.11-2)
" <Vi, Vi) <V2, Vi) • • Gi-i, Vi) <y, vx) < v<+i, vx> • • • <Vjf, V!>“
<Vi, v2> <v2, V2> • • v2> <y, v2> V2> • • • (yM> v2>
_<Vi, yM) <V2, yM) • • • <v<_!, \M) (y, vM) <▼<+!» ^M> • • •
(6.11-3)
for i = 1, 2, . . . , M.
The corresponding minimum value of the mean square deviation,
l— ||e||2, can also be determined from geometrical considerations. For
M = 2, it is the square of the magnitude of the altitude of the parallel¬
epiped determined by vl9 v2, and y. In general, it is the square of the mag¬
nitude of the altitude of the hyperparallelepiped determined by vl5 v2, . . . ,
vM, y. If Vy is used to denote the volume of this hyperparallelepiped, then
Vy = V || e ||, or
where V is the volume determined by vls v2, . . . , \M. But Vy2 is Gy, the
Gramian of v1? v2, . . . , \M, y, and V is G, the Gramian of v1? v2, . . . , yM.
This is readily apparent for the case in which the vectors determining the
hyperparallelepiped are orthogonal. For the nonorthogonal case, the
Gram-Schmidt orthogonalization procedure can be used to arrive at
the same conclusion. Thus
2
! mm min —
h° = (VTV)-1VTy (6.11-4)
VJ+1,2
V =
_VJ+1, M
Then Eq. 6.11-4 becomes
/ V y
V1
16+1 = ([VT ! V] I [VT 1 V] (6.11-5)
Lv T J -Jjr+l-
where the subscript on hj+1 indicates that J + 1 measurements are used.
Equation 6.11-5 provides a means of updating the least squares estimate
of h°. However, it is not satisfactory from a computational viewpoint if
the updating is to be continued, because of the matrix inversion required
for each new estimate.
In an attempt to avoid repeated matrix inversions, let Pj = (V2 V)_1.
Then define
V
p^+i = [Vr | v] = [Y1 V + vv7Ti-l
]
,T
so that
j+i [P^1 + vv21]-1 (6.11-6)
Direct substitution of
REFERENCES
1. J. E. Bertram, “The Concept of State in the Analysis of Discrete Time Control
Systems” 1962 Joint Automatic Control Conference, New York University, June
27-29, 1962, Paper No. 11-1.
2. D. A. Huffman, “The Synthesis of Sequential Switching Circuits,” J. Franklin
Inst., March-April 1954, pp. 161-190, 275-303.
462 State Variables and Linear Discrete Systems
Problems
6.2 Find the transfer function matrix for the system shown in Fig. P6.2.
6.3 A synchronous sequential machine accepts a serial binary coded decimal
input. After every fourth pulse, the output of the machine sends out a
signal which tells whether the last four inputs formed a correct binary
coded decimal number (0 through 9). What is the minimum number of
states of this machine?
6.4 If timing pulses tx, t2, t3, tA, tx, t2, t3, tx, . . . are available, what is the
answer to Problem 6.3?
Problems 463
Fig. P6.2
6.5 An electronic lock is designed such that, after the input sequence 101011,
the lock is opened. How many states are required in order to build this
machine?
6.6 What is a suitable set of state variables for the sampled-data system of
Fig. P6.6?
Fig. P6.6
Show that, if the matrix Ar can be diagonalized, then the periods of such
a system divide r(p — 1).
6.8 The state equation of a linear modular sequential system is given by
~3 r
x(k + 1) = x(k) mod 5
0 2
(a) Show that every state sequence has a period which divides 4.
(b) Show that there must be six sequences of period 4, covering 24 different
states. Find these sequences.
6.9 Find the minimum cycle length for the linear binary networks of Fig. P6.9.
(a)
(b)
Fig. P6.9
6.10 Set up the matrix equations for the system of Fig. P6.10 in
(a) standard A, B, C, D form
(b) normal form
Fig. P6.10
Problems 465
6.11 Using partial fractions expansion, find the matrix equations for the
following systems.
z2 + 2
(a) H(z) ^3 + ?z2 + Hz + 8 (b) H(z) =
z2 + 2z + 2
z3(z + 3)
(c) H(z) = z4 + 523 + 9z2 + 7z + 2
6.12 For the discrete system shown in Fig. P6.12, (a) which modes are con¬
trollable? (b) which modes are observable?
V2(k)
52
Fig. P6.12
V(°) / ,
T= 1
1 — e~sT
s
2
s(s + 2)
->
Fig. P6.18
6.20 For the system described by the difference equation V2?/ + 2Vy -f 2y — 1,
(a) Find the matrix formulation in terms of the standard A, B, C, D
matrices. Solve for y{k).
(b) Repeat (a) using normal form representation.
6.21 The switch in the network of Fig. P6.21 is closed at t = 0, 2, 4, . . . and
opened at t = 1, 3, 5, . . . .
7=1 ampere %
(a) Find the difference equation for the inductor current at the end of the
k\h open-close cycle.
(b) Solve the difference equation using the state transition matrix, and
draw a sketch of the inductor current as a function of k/2.
(c) Find the unit response matrix and the transmission matrix of this
circuit.
6.22 For the system shown in Fig. P6.22, (a) find the state equation formulation;
(b) solve for <!>(/:); (c) find the unit response H(&).
(c) Find the transfer function Y(z)/V(z) of the system shown in Fig.
P6.24 b.
(d) Find the unit response of this system by dividing out the transfer
function. Compare the results of this division with part (b).
6.25 (a) Find the matrix for the system of Problem 6.10 by time domain
techniques. (b) Repeat (a) using Z transform methods.
6.26 Derive the discrete analog of Eq. 5.10-14 and, from the result, the modified
adjoint discrete system.
6.27 (a) Determine a discrete approximation for the differential equation
1
y + -y = o
7.1 INTRODUCTION
+ o?y = 0 (7.2-1)
dr
x2 -f co2^ = 0 (7.2-2)
where c = constant. Thus, in any solution to Eq. 7.2-1, y(t) = xx{t) and
y(t) — x2(t) are related by Eq. 7.2-4, the equation for an ellipse. Several
solutions for various values of c, corresponding to various initial values of
xx and x2, are shown in Fig. 7.2-2. A solution in the x2 versus xx plane is
called a trajectory, and the aqa^-plane is called the phase plane.
Time is a parameter along any one of the trajectories, and for the par¬
ticular state variables chosen in this example, i.e., x2{t) — xx(t), increasing
time corresponds to clockwise motion in the aqavplane, if the axes are as
chosen in Fig. 7.2-2. As is illustrated later, however, this need not always
Fig. 7.2-1
Sec. 7.2 Phase Plane Concepts 471
y = *2
*2 = C 3
Fig. 7.2-2
(7.2-5)
For example, let point a be defined by (0, c3) and point b by (c3/co, 0) on the
c — c3 trajectory. Thus a and b are one-fourth of a period apart. Elimina¬
tion of x2 from Eq. 7.2-5 by substitution of Eq. 7.2-4 for c = c3 yields
(7.2-6)
trajectories were paths continuously diverging from the origin, one would
intuitively say that the system is “unstable.” This representation of the
behavior of a second order system and indication of “system stability”
without actually solving the system differential equation is an important
use of the phase plane approach in connection with nonlinear systems. It
is necessary, however, to provide more precise definitions of stability. This
is done in Section 7.7.
The center singular point observed for the linear oscillator is only one of
four possible singular points. The remaining types of singular points can
be illustrated by means of the system of Fig. 7.3-1. Notice that, for the case
of zero damping (| = 0), this system is the same as the linear oscillator.
Thus, for £ = 0, the origin of the aqa^-plane is a center, and the trajectories
are ellipses. For ^ 0, however, the trajectories are altered significantly.
The differential equation for the system of Fig. 7.3-1 is
(7.3-2)
x2 = — c OiX1 — 2^ cox 2
Fig. 7.3-1
Sec. 7.3 Singular Points of Linear, Second Order Systems 473
(7.3-4)
s2 = —£<jj + coV £2 — 1
Thus at this point one could plot the response y{t) for various values of £
and co, since the general solution of Eq. 7.3-1 is
[cox-l + 2£x2]
(7.3-5)
m — — — — (7.3-6)
x1 xx
Since Eqs. 7.3-7 and 7.3-3 have the same roots, the slopes of any straight-
line trajectories are equal to the characteristic values given by Eq. 7.3-4.
This result is, of course, dependent upon the choice of the state variables,
although the characteristic values themselves are independent of the choice
of state variables. These straight-line trajectories correspond to the modes
of the response, and could have been obtained from Eq. 5.6-10.
With these preliminary thoughts completed, the nature of the singular
points for linear, second order systems is now considered.
This case was investigated in the preceding section. It was found that the
origin is a center, and that the trajectories are ellipses about the origin.
No straight-line trajectories exist, since there are no real roots to Eq. 7.3-7
for £ = 0, i.e., there are no modes or characteristic vectors corresponding
to real characteristic values.
For this case, it is not possible to carry out directly the procedure of the
previous section to determine the equation for the trajectories. However,
it is again possible to say that there are no straight line trajectories, since
Eq. 7.3-7 has no real roots for |£| < 1.
The substitution of x2 — zxx and dx2 = z dx1 + xx dz permits Eq. 7.3-5
to be written as
dxx z dz
(7.3-8)
z“ -f- 2h,coz -}- oo
Sec. 7.3 Singular Points of Linear, Second Order Systems 475
21 x2 + £wx1
(x2 -j- £ojx1)2 + co2( 1 — £2)xf = c2 exp tan -i
U/i - ? (;
Wi -
2f
_
- tan 1 Z2
2 . 2 2
z2 + % = c exp (7.3-9)
Wl - I2 2 1-1
upon introducing the coordinates
£oj 1 Xc
z1 — r cos (/>
z2 = r sin </>
£<f>
r = c exp —
Wl - f2J
Thus the trajectories in the z^-plane are logarithmic spirals. If 1 > £ > 0,
the radius decreases as <f> becomes more negative (which is the direction of
increasing t), corresponding to Fig. 1.3-2a. This is the case of a stable focal
point. The unstable focal point corresponding to 0 > £ > — 1 is illus¬
trated in Fig. 1.3-2b. As |£| is increased from zero toward unity, the rate
at which the radius changes with <f is increased. This corresponds to
moving the system poles, as given by Eq. 7.3-4, away from the joo axis of
Fig. 7.3-2
476 Introduction to Stability Theory and Lyapunov's Second Method
the 5-plane and agrees with the result one would intuitively expect. The
trajectories about the focal point in the z^-plane are distorted versions
of Fig. 7.3-2 and are illustrated in Fig. 7.3-8. Example 5.6-2 is a specific
case of a stable focal point.
In the situation in which the origin is a nodal point, there are two trajec¬
tories which are straight lines. The slopes of these trajectories are given
by the roots of Eq. 7.3-7 and are the characteristic values of Eq. 7.3-4.
These two trajectories are illustrated in Fig. 7.3-3 for £ > 1. The remain¬
ing trajectories about a nodal point in the aqa^-plane are more difficult to
evaluate.
In order to determine the remaining trajectories, it is useful to introduce
the transformation z = M_1x or, more specifically,
«1
(7.3-10)
_ x 2_
The zx- and z2-axes are the straight-line trajectories shown in Fig. 7.3-3.
Thus the z1z2-plane is a distortion of the xxx2-plane. Using the transforma¬
tion of Eq. 7.3-10 and Eq. 7.3-4, Eq. 7.3-2 becomes, for the case of a nodal
point, the normal form equations
«1 = Sfa
(7.3-11)
z2 = SXZ2
dz2 _ s1 dz1
z2 s2 zx
x2
Fig. 7.3-3
Sec. 7.3 Singular Points of Linear, Second Order Systems 477
If the sign of the gain in the outside feedback loop of the system in
Fig. 7.3-1 is changed, the singular point becomes what is known as a
saddle point. For this case, the characteristic equation is
s2 + 2£ | co | s — | co21 = 0
The corresponding characteristic values are
*2
Fig. 7.3-5
Summary
The conditions on Eq. 7.3-1 for the various types of singular points are
summarized in Fig. 7.3-7. Not considered are the case for which ||| = 1
and the case for which oo2 = 0. The trajectories for these cases are
contained in Fig. 7.3-8. The corresponding expressions for the trajectories
are given in the problems.
22 22
Fig. 7.3-6
Sec. 7.4 Variational Equations 479
The time response of the linear, second order system is easily determined,
so that it is possible to compute the phase plane trajectories from the time
response. For example, Eqs. 5.6-8 and 5.6-14 could have been utilized.
This was not done in this section, because it is not generally possible for
nonlinear systems, and the interest in studying the phase plane trajectories
for linear systems stems from their relationship to nonlinear cases. In non¬
linear cases, past studies often resorted to graphical methods of deter¬
mining the trajectories, such as the method of isoclines, Lienard’s method,
and the phase plane delta method.2-6 In the light of present-day computer
simulation capabilities, however, such graphical techniques lose much of
their appeal. Except for a brief discussion of the method of isoclines in the
next section, the graphical techniques are not pursued further in this book.
This is because of the introductory nature of this chapter and because of
the belief that the proper use of the phase plane method, in a case where the
trajectories cannot be determined analytically, is in conjunction with a
computer simulation. The trajectories can be measured from the simula¬
tion, and the phase plane approach enables one intelligently to determine
proper parameter changes to improve system behavior. For the same rea¬
sons, some of the more sophisticated methods for determining the values
of time along a trajectory are not considered.7
y + 2£coy + co2y = 0
i vX2
> 'Sl
Center 'X \ - T
£ = 0. co2 >0
1
> ' S2
V
_ vX2
X S2
Stable focus
»—7—j-Xi
0<£ <1, co2>0
XS2
XS2
Unstable focus
(
4
\ > Ti
-1 <£<0, co2>0
X si
\ JJ
\ t .*2 \ \
\ = S'2Xi
\X2 = SiXi
/ X9 = S9Xi
\x2 /
/ / ‘
r -/= sixi
Unstable node
-X—X—
£ < -1, co2> 0 S1 S2
Fig. 7.3-8
Sec. 7.4 Variational Equations 481
y + 2£ooy + co2y = 0
X9 = S9Xl
Saddle point
J
v v
co2<0 A A 1 ^ X
(figure is for £ = 0) si S2
n *2 = SiXi
U2 \ \
Si, S2
S= l V 1 I > X1
stable
X2 = —OOXl
X2 = COXl
/ / i
si, S2
f=-l
unstable a ) 1 > *1
N. N. )
[X2 n\^X2 = — 2£coxi + c
co2 = 0 /
X ?\ \ N. > *1
> 0 -2£w
X'
^2 / ^
GO2 = 0
?/\ V
X A A
£w <0 2£w
y^X2 = 2£wxi + c
This is the autonomous case, that is, one in which the independent variable
t does not appear explicitly. It corresponds to a system which is both
unforced and time-invariant. Equation 7.4-1 is an abbreviated representa¬
tion of
X1 =f1(x1, x2, . . . , xn)
Thus x and f(x) are /7-dimensional vectors. The singular points of Eq.
7.4-1 are given by x = 0. Hence they are the solutions of
f(x) = 0 (7.4-2)
In the case of a linear system, f(x) is a linear function of x. Thus Eq. 7.4-2
has only the solution x = 0, provided that the determinant of the coeffi¬
cient matrix of the linear system is nonzero. This is true if there are no
free integrators. Then only one singular point exists for the corresponding
linear system, and it is at the origin of the state space. However, in the
nonlinear case, more than one singular point can exist, as is evidenced by
the possibility of more than one solution to Eq. 7.4-2. Denoting the zth
solution by xie, Eq. 7.4-2 becomes f(xie) = 0, z = 1,2,.... If each of the
components of f(x) can be expanded in a Taylor series about the zth
singular point, the result of considering only the linear terms is
U = (7.4-5)
The question then arises as to how the stability of the system of Eq. 7.4-5
relates to the stability of Eq. 7.4-1.
It is possible to state that, if any of the roots of Eq. 7.4-6 has positive
real parts, then the zth singular point is unstable. If all the roots of Eq.
7.4-6 have negative real parts, then the zth singular point is stable.8 (In
this case the singular point is actually what is known as asymptotically
stable. However, asymptotic stability is not defined until Section 7.7.)
These statements about the stable and unstable behavior of the singular
points of a nonlinear system are given here without proof. They are
most easily proved by using Lyapunov’s method, which is introduced in
Section 7.8.
For the case in which the variational equations have roots with zero real
parts, it is impossible to distinguish between stability and instability of the
singular point based on the linear approximation.9 One can intuitively
view this situation as a borderline case, in which the effect of the ignored
nonlinear terms can result in either stable or unstable behavior.
Although the preceding statements concerning the stability of the sin¬
gular points of a nonlinear system, as indicated by the variational equations,
are valid regardless of the order of the system, a second order example is
chosen so that a phase plane may be used to illustrate the results.
Example 7.4-1. Figure 7.4-1 is a simplified pictorial representation of a relaxation
circuit.10 Its differential equation is
dvr d2i di
-Keq = - KRC-± = - KRCL — = L — + RLi + v (7.4-7)
9 dt dt2 ,dt
Fig. 7.4-1
b2 (7.4-10)
x2 — — xx — c2xx3 — 2 ax2
Equating the right side of Eq. 7.4-10 to zero determines that there are three singular
points located at
(1) xx = 0 (2) xx = b/cV2 (3) xx = -b/cVl
x, = 0 X, 0 Zo = 0
The Jacobian matrix of Eq. 7.4-4 evaluated at the first singular point is
"0 1 '
Ji =
p2l 2 -2 a
X2 = i
neighborhood of its singular points is as shown in Fig. 7.4-3. Notice that this does not
indicate the system stability or response for an arbitrary initial condition, but merely
indicates the behavior in a small region about each of the singular points. In fact, the
trajectories as illustrated are not precisely correct, since the trajectories about a given
singular point will be distorted from the curves determined for the linear case by
neighboring singular points.
The determination of the trajectories in larger regions of the phase plane is not an
easy task. One approach is to use the method of isoclines, which consists in determining
curves (for linear systems, they are straight lines) which connect points in the phase
plane where the trajectory slopes are equal. One can then sketch through these points
with the appropriate slope and estimate the trajectories.
This procedure is illustrated for this example by using Eq. 7.4-10 to write
Equation 7.4-11 gives the slope at any point in the phase plane. Assuming that the slope
is a constant k, Eq. 7.4.11 becomes
If c were zero (corresponding to a linear system), Eq. 7.4-12 would be the equation of a
straight line, and the isoclines could easily be evaluated. For this case, however, it is
necessary to pick a value for k and then compute the value of x2 for various values of xx.
The results are the isoclines shown in Fig. 7.4-4 for the arbitrarily selected case of a =
c = 1, b = 2. Sketching through each of the isoclines with the appropriate slope permits
the trajectories of Fig. 7.4-5 to be determined.
486 Introduction to Stability Theory and Lyapunov's Second Method
Fig. 7.4-4
Sec. 7.4 Variational Equations 487
II
K
Separatrix
488 Introduction to Stability Theory and Lyapunov's Second Method
The trajectories of Fig. 7.4-5 are useful in that they indicate the system behavior for
any of the initial conditions considered therein. They show two stable conditions.
Thus the system could be used as a flip-flop. However, for the parameters chosen, the
design is poor in the sense that an extremely large triggering signal is required to change
the system from one stable condition to the other. This is revealed by the fact that the
triggering signal would have to drive the system states across the curve labeled separa-
trix. A separatrix is a trajectory which passes through singular points and divides the
phase plane into regions of different character to the trajectories. The separatrix of Fig.
7.4-5 separates the trajectories about the left and right focal points. Since the negative
characteristic value of the saddle point is approximately the slope of the separatrix,
the phase plane analysis reveals that the size of the triggering signal can be altered by
changing a. Adjustment of b would also change th£ triggering signal requirements,
but it has the simultaneous effect of changing the separation of the singular points.
approach it, while those on the other side of the limit cycle leave it. Such
a case is shown in Fig. 7.5-lc.
The existence or nonexistence of limit cycles in the behavior of a system
is of fundamental importance to engineers. For example, a control system
engineer generally desires systems without limit cycles, although small-
amplitude oscillations are sometimes acceptable. On the other hand, an
engineer designing an oscillator would definitely want a system with a
stable limit cycle. Several theorems are available to guide the engineer in
this respect.
Poincare’s Index1213
The index n of a closed curve in the phase plane is given by N, the total
number of centers, foci, and nodes enclosed, minus S, the number of
saddle points enclosed. That is, n = N — S. A necessary condition for a
490 Introduction to Stability Theory and Lyapunov's Second Method
closed curve to be a limit cycle is that its index be +1. This criterion is not
sufficient, however. As an illustration, consider Example 7.4-1. A large
closed curve enclosing the two focal points and the saddle point in Fig.
7.4-5 has an index of +1. Thus a limit cycle conceivably could exist which
encloses all the singular points. This is not the case, however, as can be
shown by Bendixson’s negative criterion.
This criterion is sometimes useful for proving that no limit cycle exists
in a region of the phase plane. Consider tile equations xx — fx{x^ x2),
x2 = f2(x1, x2). The slope of any trajectory is given by
dx 1 xx /iOj, x2)
This can be rewritten as fx(xl5 x2) dx2 — f2(x1, x2) dxx = 0. Thus, around a
limit cycle,
does not change sign or vanish identically within a region of the phase
plane, the integral of Eq. 7.5-2 cannot be zero. Since Eq. 7.5-2 applies
along a limit cycle, no limit cycle can exist within a region of the phase
plane in which / does not change sign or vanish identically.
Example 7.5-1. As an illustration of the use of this theorem, determine I for Example
7.4-1.
Use of Eqs. 7.4-10 and 7.5-3 yields I = —2a. Since this does net change sign, nor is it
zero anywhere in the aq:z2-plane, no limit cycle can exist for the flip-flop of Fig. 7.4-1.
Thus the circuit of Fig. 7.4-1 cannot oscillate. This is an important property for a flip-
flop.
x1 = x2
Fig. 7.5-2
Assuming 0 < £ < 1, this system has only one singular point, an unstable focus, at
x-l = x2 = 0. Thus its index is -fi 1 and a limit cycle could exist. Apply Bendixon’s
negative criterion.
Evaluation of I given by Eq. 7.5-3 yields
/ = 2|
a
Since I does not change sign or identically vanish for \x2\ < Va, no limit cycle can exist
which is wholly contained within the region specified by \x2\ < Va, assuming a > 0.
A limit cycle which is not wholly contained within this region does exist, however, since
/
Eq. 7.5-4 corresponds to the Rayleigh equation14 y — 2$y II — — I
V \ y — 0. Notice,
from the expression for /, that no limit cycle can exist for negative values of a.
Poincare-Bendixson Theorem15
*2
Let Cx be a circle of radius r with its center at the origin, so that xx2 -f x22 = r2. Then
the slope of the circle is given by
dx 2 xx xx
dxx Vr2 — xx2 X2
The difference between the slope of the trajectory and the slope of the circle at a common
point is
For all |aj2| < V3a, 6 > 0, so that the slope of the trajectory is more positive than the
slope of the circle. This indicates that all trajectories pass out of Cx for any r < V 3a.
Thus the system of Fig. 7.5-2 has no limit cycles within the region defined by r < V3a.
This region is larger than that defined by Bendixson’s negative criterion.
In this example, it is not possible to use similar reasoning to determine a circle C2 for
which all trajectories enter, in an attempt to locate the limit cycle. The necessary curve
is more complicated.16
Fig. 7.5-4
m(a) = 0, Region I
m(a) = — 1, Region II
/77(a) = -f 1, Region III
and the trajectories in each of these regions correspond to those of a linear system. Use
the successor function to determine the conditions for a limit cycle.
The trajectories in each of the three regions can be determined from the state equations
x1 = x2 and x2 = —x2 + m(a). The trajectory slopes are given by
(b)
Fig. 7.5-5
Fig. 7.5-6
Sec. 7.5 Limit Cycles 495
This can be integrated directly to give the expression for the trajectories in Region I.
They are the straight lines
shown in Fig. 7.5-6. It should be observed that there is an infinity of singular points in
Region I given by la^l < d, x2 = 0.
In Region II, Eq. 7.5-5 is
dx 2 1 + x2
(7.5-7)
dx x x2
Integration of Eq. 7.5-7 yields
The trajectories in Region II are shown in Fig. 7.5-6. They are shown only for x2 > — 1,
since Eq. 7.5-8 is valid only in this range. Since the slope of the trajectories is zero at
x2 = — 1, as indicated by Eq. 7.5-7, no trajectories can cross x2 = —1. Thus no limit
cycle can exist outside the range [rc2| < 1. The trajectories in Region III have the same
shape as those in Region II, but they are rotated by 180°.
The three regions in Fig. 7.5-6 are divided by the four switching lines as indicated,
assuming k > 0. These lines are the locus of points at which the relay output switches
value. The reader should coordinate these switching lines with the relay characteristic
of Fig. 7.5-5b and verify that such is the case.
A possible stable limit cycle is shown in Fig. 7.5-6. It is yet to be determined if a limit
cycle actually exists. In order to do this, a successor function will be determined.
Because of the symmetry which exists in this example, however, the successor function
need span only one-half of the possible limit cycle. That is, a successor function will be
determined which relates the point [x10 = — {hd + kx20), x20] to the point [x12 = hd —
kx22, x22]. If x12 = —x10 and x22 = — x20, then a limit cycle exists.
From Eq 7.5-6, it is apparent that the points (x10,x20) and (xxx, x2X) are related by
d{ 1 -f- h)
#21 = #20 - -7- (7.5-10)
1 — k
Similarly, Eq. 7.5-8 and the appropriate switching line expressions can be used to relate
x2X and x22 by
d( 1 + h)
(1 + x22) exp [—^(l — k)] = 1 + #20 exp [— a?20(l — k) + 2dh]
1 - k
(7.5-12)
Although this expression cannot be solved explicitly for x22 in terms of x20, it is in essence
a successor function relating these two coordinates. If x22 = — x20 satisfies Eq. 7.5-12,
496 Introduction to Stability Theory and Lyapunov s Second Method
Fig. 7.5-7
there is a limit cycle. This is true since this equality would also necessitate x12 = —x10,
because the two points are on lines of equal slope, symmetrically located about the
origin.
In order to determine if any solutions to Eq. 7.5-12 and x22 = —x20 exist, let
d( 1 + h)
p0(x) = 1 + x — exp [—^(l — k) + 2dh\
1 - k
and p2(x) = (1 — x) exp [x{\ — k)]. If there is a value of x for which p0(x) = p2(x),
then that value of x is the x20 = — x22 satisfying Eq. 7.5-12, since p0(x) and p2(~x)
correspond to the right and left sides of this equation, respectively. Typical curves of
p0(x) and p2(x) with an intersection, and hence a limit cycle, are shown in Fig. 7.5-7
for x in the range d( 1 + h)/( 1 — k) < x < 1. The upper limit was specified previously
in connection with Eq. 7.5-8. The lower limit is a consequence of the fact that the Region
I portion of any limit cycle trajectory cannot intersect the region of singular points,
otherwise the system would come to rest. The initial conditions for which no periodic
motions exist are shown in Fig. 7.5-9 in Area I.
From the values of p0(x) and p2{x) at x — <7(1 -f h)/( 1 — k), it is apparent that the
curves do not intersect unless
d( 1 + h)
e
—2d
< (7.5-13)
1 - k
Thus Eq. 7.5-13 is a necessary condition for a limit cycle. The values of dand h satisfying
Eq. 7.5-13 for various values of k are those less than the values on the curves of Fig.
7.5- 8. For k > 0.5, no limit cycle can exist.
In order to show that a limit cycle exists if Eq. 7.5-13 is satisfied, assume that x20 —
x = xa in Fig. 7.5-7. Then p0(xa) = Pztxb) corresponds to a solution of Eq. 7.5-12, so
that the resulting x22 = — xb. This is the value of x20 for the next half-cycle, so that the
new x22 = xc. This procedure continues until x20 = x22 = x0. This is a stable limit
cycle, since the same result is achieved starting from xa'.
For the case illustrated, the phase plane can be divided into three areas as in Fig.
7.5- 9. If the initial conditions are such that the initial point is in Area I, the system is
Sec. 7.5 Limit Cycles 497
Fig. 7.5-8
498 Introduction to Stability Theory and Lyapunov's Second Method
stable and the trajectory goes to the singular region. If the initial point is in Area II,
the trajectory approaches the stable limit cycle from within. If the initial point is in
Area III, the trajectory approaches the stable limit cycle from outside.
This example has demonstrated the use of the successor function for determining
conditions under which limit cycles can or cannot exist. Typical results are those of
Fig. 7.5-8, which indicate the amount of rate feedback, in terms of the relay deadband
and hysteresis, required to prevent the existence of a limit cycle. Figure 7.5-7 illustrates
a method of determining the size of the limit cycle if one exists.
Summary
This section has considered limit cycles, and some of the theorems and
methods which may be useful in determining their existence. They are, in
essence, methods of estimating if a system has periodic motion without
explicitly evaluating the system response. Unfortunately, the phase plane
approach is limited primarily to unforced, second order systems. Theoreti¬
cally at least, the concepts may be extended to higher order phase space to
treat systems of higher than second order.5,19 In general, however, the
extension has not been an overwhelming success. This is primarily due to
the problem of determining, presenting, and visualizing trajectories in
phase space. For this reason, most of the theory has not been extended to
include higher order systems.
(7.6-1)
x2 = —k(x 1)
Assuming /:(0) is zero and /c(aq) 3^ 0 for aq ^ 0, there is only one equilib¬
rium point, a center, located at the origin of the aqaq-plane. The trajec¬
tories enclose the center and are given by the solutions of the equation
dxr /c(aq)
(7.6-2)
daq x2
Equation 7.6-2 can be integrated by separation of variables to yield
Xo
+ /c(aq) dxx = c, c = constant (7.6-3)
Thus the trajectories take the form shown in Fig. 7.2-2. The exact shape
depends upon k(y).
The total energy of this conservative system is the kinetic aq2/2, plus the
potential
/c(aq) dxx
Equations 7.6-3 and 7.6-4 yield E(x1? x2) = c. In other words, the xxx2
trajectories are contours of constant total energy for this conservative
system. Alternatively, the time rate of change of the system energy is zero,
as is also shown by
dE
— = x2x2 -f k(x1)x1 = x2[x2 + ^(aq)] = 0 (7.6-5)
dt
Fig. 7.6-2
Sec. 7.7 Stability and Asymptotic Stability 501
Substitution of Eq. 7.7-2 into Eq. 7.7-3 yields the differential equation for
the perturbed motion x(t) as
Equation 7.7-4 has an equilibrium point at the origin, since it has the
trivial solution x = 0. Thus one can now consider stability of the origin
of the equivalent system
x = f(x, t) f(0, 0 = 0 (7.7-5)
where f(x, t) = g[x + us, t] — g[us, t]. Unfortunately, Eq. 7.7-5 is often
more complicated than the original version, Eq. 7.7-1. Furthermore, the
specific solution of Eq. 7.7-1 is required to determine Eq. 7.7-5. However,
Eq. 7.7-5 does permit the definitions of stability to be formulated in terms
of the stability of an equilibrium state at the origin, and most of the
Sec. 7.7 Stability and Asymptotic Stability 503
q 12 ... qx
qn o,
7n q\2
>0, .. ., ^21 q 22 ... q2
^21 ^22
qn i qn2 • • • qn
This theorem was proved in Section 4.9.
,,, . dV dV dV , dV .
H-x n (7.8-2)
dt dXl dx2 dx n
i\ = x2, . . . , xf)
x2 — x2-> • • • J Xn)
dV
w= y-Aw + dV + + dx
p-fjx)
U U OC'2 n
In terms of the gradient of F(x), denoted by grad V(x), the time derivative
of V(x) becomes
W = (grad F, f> (7.8-3)
The function W is the time derivative of F(x) for the differential equation
of Eqs. 7.8-1. Note, however, that the solutions of Eqs. 7.8-1 are not
required to evaluate W. v-
Lyapunov’s stability theorem can be written in terms of the functions
F(x) and W.
x 2 CXl
E(xlt x2) = V{xx, x2) = — + k(u) du (7.8-4)
2 Jo
Note that F(0) = 0, and F(x) is positive definite since any physical spring
is such that k(u) has the same sign as u. From Eq. 7.6-5, the time derivative
of F(x) is
dE
=~= W = x2[xt + kizj] (7.8-5)
dt
Fig. 7.8-1
The state equations are x1 = x2 and x2 = —gixj — ax2, where it is assumed that
g(xt) = 0, only at xx = 0, so that the only equilibrium is at xx = x2 = 0. From the
similarity of these equations to those of Eqs. 7.6-6,
x 2 f*i
V(x) = -L + g(u) du
1 Jo
is suggested as a possible Lyapunov function. Note that K(0) = 0, and V(x) for x#0
is positive if the^p/) versus u characteristic is anywhere in the first and third quadrants
as shown by the solid curve in Fig. 7.8-2. The latter condition can be expressed as
g(u)
7— >0, u 0 (7.8-6)
u
g(u)
Fig. 7.8-2
Example 7.8-8.32,z3 Barbasin investigated the stability of the third order nonlinear
differential equation
d3y d2y
+ a + h{y) = 0 (7.8-7)
dt» dt2
xx = x2
x2 — xz (7.8-8)
Fig. 7.8-3
Sec. 7.8 Lyapunov s Second Method for Autonomous Continuous Systems 509
'*2
V(x) = a h(r) dr + g(r) dr + %(ax2 + x3)2 + x2h(xx)
g(x2) dh(xi)
W— — a x0
x2 dx1
Then, using reasoning similar to that of Example 7.8-7, the equilibrium of the system of
Fig. 7.8-3 is asymptotically stable if V(x) is positive definite. These requirements on
V(x) are now examined.
If a and h{u) are such that
a > 0
h(u)
— > 0, u 0 (7.8-10)
u
then the conditions for positive definiteness of the first term of the T-function are
satisfied. Furthermore, satisfaction of Eq. 7.8-10, in conjunction with Eq. 7.8-9, yields
g(x2)jx2 > 0 for x2 0. This means that the second term of the K-function is positive
definite if Eqs. 7.8-9 and 7.8-10 are satisfied. Also, the third term of the E-function
satisfies the positive definite conditions, with the exception that it is zero for ax2 = —x3.
Even if this is the case, however, V(x) is not zero since the second term is positive.
Thus it remains to consider the last term in the E-function for x2 ^ o. When x2 ^ 0,
it is possible to write the E-function in the form
x-\ r x<>
g(r) dh(u)
[2 G(x2) + x2h(x i)]2 + 4 h(u) a r dr du
fo Jo du
L(x) = Max 2 + x3)2 +
4 G{x2)
where
‘X2
G(x2) = g(r) dr
The sum of the first term of E(x) and the first term of the fraction in E(x) is always
positive for nonzero x2. Also, the integration with respect to r is positive because of
Eq. 7.8-9. Hence the integration with respect to u is positive. Thus it is now possible to
say that E(x) is positive definite and the equilibrium is asymptotically stable if Eqs.
7.8-9 and 7.8-10 are satisfied.
It is interesting to note that the criteria for asymptotic stability of the equilibrium
contain two different types of linearizations.23,34 These are dh{u)\du and h(u)/u. These
linearizations are represented in Fig. 7.8-4 at a point u = u0. The first linearization is
the slope of the function, while the second is the slope of the intercept. If h(u) were
voltage and u were current for a nonlinear resistance, then dh/du would be the a-c or
dynamic resistance, while h(u)/u would be the d-c or static resistance, both evaluated
at
If the dynamic and static linearizations are the same, as they would be for a linear
h(y) function, and if g(y) is linear, then it is possible to replace h(y) by a0y and^O/) by
axy in Eq. 7.8-7. It then becomes
d*y d2y dy
(7.8-11)
d? + aiP + a'dt + a°y = 0
a > 0
a0>0 (7.8-12)
aax — a0 > 0
These are precisely the Routh-Hurwitz conditions for stability of the linear system
described by Eq. 7.8-11. Unfortunately, however, when one attempts to analyze a
nonlinear system by introducing some type of linearization, it is usually not evident
which type of linearization should be used. This is illustrated by this example, in which
both types of linearizations are involved.
Suppose that k(u) is such that the potential energy term does not become
infinite as x1 approaches infinity, but approaches a finite limit c0. If
c0 < c, the curves of constant V may not all be closed curves as in Fig.
7.6-2, but they could appear as in Fig. 7.9-1. In this case the origin would
be asymptotically stable, but not asymptotically stable in the large. It
would not be asymptotically stable in the large, since for sufficiently large
x0 it is possible for the system to move from a state of higher energy to a
state of lower energy without approaching the equilibrium at the origin.
The requirement for asymptotic stability in the large removes this
possibility.37
Example 7.9-1. Determine the conditions for which the equilibrium of the system of
Fig. 7.8-1 is asymptotically stable in the large.
In this case, Eq. 7.8-6 must be replaced by g(u)/u > a > 0, // ^ 0. a is a constant.
This ensures that F(x) approaches infinity as ||x|| approaches infinity.
Example 7.9-2. Determine the conditions for which the equilibrium of the system of
Fig. 7.8-3 is asymptotically stable in the large.
512 Introduction to Stability Theory and Lyapunov's Second Method
X2
g(x 2) dhM
—-— > aoc2 — ax > 0, 0
a- X2
x2 dx1
are the equations of motion of a rigid body written about the principal axes of inertia.
Ix, Iy, and Iz denote the moments of inertia about these axes; co*, cov, and coz, the angular
velocities; and Mx, Mv, and Mz, the external torques.
Assuming that the rigid body is a space vehicle tumbling in orbit, it is desired to stop
the tumbling by applying control torques proportional to the angular velocities. The
control torques are Mx = —kxojx, My — —kycov, and Mz = —kzcoz. Lyapunov’s
method can be used to determine the stability of the responses.
Choosing the state variables xx = cox, x2 — tov, and x3 = coz permits the system to be
described by
x = A(x)x (7.9-2)
where
A(x) = (7.9-3)
Sec. 7.10 First Canonic Form of Lur'e—Absolute Stability 513
Note that the system has an equilibrium at x = 0. Let F(x) be the positive definite
quadratic form of Example 7.8-6, i.e., F(x) = <x, Qx>, and choose
0 0'
Q = o If 0 (7.9-4)
0 o If
so that F(x) = Ifxf -f Ifxf + Ifxf, the square of the norm of the total angular
momentum. The corresponding time derivative of F(x) is W = (x, Qx) + <x, Qx).
Using Eq. 7.9-2 yields W = (A(x)x, Qx) + (x, QA(x)x). This can be rewritten as
Thus W is negative definite if P is positive definite. Then, since F(x) is positive definite,
and because of its form must approach infinity as x approaches infinity, the equilibrium
is asymptotically stable in the large, if P is positive definite. Substitution of Eqs. 7.9-3
and 7.9-4 into Eq. 7.9-6 yields
2 kxIx 0 0 ■
P = 0 2 kyly 0
0 o 2 kzIz
If each of the k's is positive, P is positive definite and the equilibrium is asymptotically
stable in the large.
t The case in which some of the A’s occur in complex conjugate pairs with negative real
parts can be handled in a somewhat similar fashion.
514 Introduction to Stability Theory and Lyapunov's Second Method
(a)
(b)
Sec. 7.10 First Canonic Form of Lur'e—Absolute Stability 515
where the scalar r — — (b0 + bx + • • • + bn) and the vector X has com¬
ponents 2l9 2.2, ..., 2n.
A Lyapunov function for this system is
where Q is positive definite and k(a) is restricted to the class for which
+ k(a)[(k, x) - rk(a)]
IF is a quadratic form in x and k(a) and is negative definite if and only if44
<3n • • • 0
22
qu > 0, i = 1, 2, ..., n
0 <7 nn
where all the nondiagonal elements are zero. Then Eq. 7.10-4 becomes
r > —S v (7.10-5)
where S is given by
s _1 y Ibj + W8
4z=i A-i/qa
— b0 > bx + b2 + • • * + bm (7.10-6)
It is readily apparent that Eq. 7.10-6 is applicable only if (—b0) is more
positive than the sum of the positive residues of G(^). Since Eq. 7.10-3,
together with Eq. 7.10-6, form sufficient, but not necessary, conditions for
asymptotic stability in the large, failure to satisfy these conditions does not
necessarily mean that the system is unstable. It may mean that the
criterion does not apply in the particular case under consideration.
If Eq. 7.10-6 is satisfied, the equilibrium is asymptotically stable in the
large for any nonlinear characteristic satisfying Eq. 7.10-3. Stability of
this type, i.e., asymptotic stability in the large for somewhat arbitrary Ar(cr),
is called absolute stability. Various other absolute stability criteria can be
derived by modifying the previous procedure (e.g., by choosing a different
Q), or the Lyapunov function.45,46
Example 7.10-1. The absolute stability of the equilibrium of the system with G(s) =
—(5 + 3)/(s + 2)(j — 1) is to be investigated.
Since G(s) does not have a pole at the origin, Eq. 7.10-6 is of no direct use. However
it can be utilized by making use of a technique known as pole-shifting.47 The pole-
shifting technique can be illustrated by means of Fig. 7.10-2. The feedforward and feed¬
back f's cancel so that the basic system and its conditions for stability remain unchanged.
Sec. 7.10 First Canonic Form of Lur'e—Absolute Stability 517
Fig. 7.10-2
However, the feedback around G{s) can be utilized to move the poles of G'(s), since
G(s)
G'(s) =
1 - <f>G(s)
In the example under consideration,
~(s + 3)
G'(s)
j2 + (1 + <f>)s + (3<f) - 2)
so that choosing <f> = -I results in G'(s) having a pole at the origin. Now Eq. 7.10-6 can
be applied.
For <£ = f, 9 4
—(i1 + 3) .
G\s) = i + _^
s(s + f) 5 s + f
k\o) 'k(o)—%o
> e > 0
a
Since k\o) is restricted to the first and third quadrants by Eq. 7.10-3, k(o) must be
restricted to the unshaded area of Fig. 7.10-3. If k(o) is replaced by a linear gain k, the
Routh-Hurwitz condition on k for stability is precisely that of Fig. 7.10-3. Aizerman
has shown that this is true in general for second order systems.48
k(o)
518 Introduction to Stability Theory and Lyapunov's Second Method
Theorem 7.11-1. The equilibrium of Eq. 7.11-1 is totally stable, if for every
e > 0 there is a (5(e) > 0 and a ^(e) > 0, such that, if ||x0|| < S and
|| u(x, t) || < (5X for all x and t > t0, then || x(t; x0, t0) || < e for all t > t0.49-53
Theorem 7.11-2.53 Let L(x) be a scalar function which for all x has contin¬
uous first partial derivatives and be such that V(x) approaches infinity as
|| x || approaches infinity. If W, the time derivative of V(x), evaluated for
Sec. 7.12 Estimation of Transient Response 519
Eq. 7.11-2 is such that W < — e < 0 for all x outside R0, for all u(x, t) in
U, and all t > t0, and if V(x) < K(q) for all x in R0 and all q outside R,
then Eq. 7.11-1 possesses a strong practical stability.
ATx kxcox T
My ~ —kyCOy -f Uy
where the control torques are as previously considered, and the disturbance torques on
the x, y, and z axes are ux, uy, uz, respectively.
Using the same F(x) function as in Example 7.9-3, W for the perturbed system is
where uu u2, and u3 have been written for ux, uy, and uz, respectively. Let ulm, u2m, and
u3m denote the largest absolute values for the disturbances in the region U. Choosing
the region R0 to be the interior of an ellipsoidal surface containing the three points
(,U\mlkx, 0, 0), (0, u2Jky, 0), and (0, 0, u3Jkz), then W < — e < 0 for all x outside R0,
all u(x, t) in U, and all t > t0. Also, by choosing R to be some region larger than R0,
V(x) < V(q) for all x in R0 and all q outside R. Thus the system possesses a strong
practical stability. Note that, if the maximum disturbances are increased, the same
regions R0 and R can be maintained by increasing the gains kx, kv, and kz.
Fig. 7.12-1
520 Introduction to Stability Theory and Lyapunov's Second Method
dVjdt _ _ W _ 2
(7.12-1)
V ~ V ~ T
W 2a
— 2 7~ 0
V fxi
1 + (2[x22) g(u) du
Jo
Discuss the effects of the system parameters on the speed of response.
From the viewpoint of speed of response, the designer should choose a as large as
possible. One must be careful about drawing conclusions concerning the functional )
however, since its effect on — W/V varies with the state variables x1 and x2.
The concept of the previous section can be used for the design of relay
controllers for controlled elements described by
x = Ax + Bm (7.13-1)
'a2 + b( 1 + b) 1
lab 2b
Q =
1 1 + b
2b lab
B^Qx can be determined to be
0
BQx = xi (1 + b)x 2
2b + lab
Thus mx — 0 and
(1 + b)x2(t)'
m2(t) = —M2 sgn xi (0 + a
where the positive constant 1/2b has been removed from each of the terms within the
brackets, since it does not affect the sign of m2(t).
The resultant control system is asymptotically stable in the large by design. Further¬
more, if u is the maximum value of a disturbance at the point indicated in Fig. 7.13-1,
then the system has a strong practical stability since, for xx = x2 and x2 = —bxx —
ax2 + m2 + u, IF is given by
x1 (1 + b
W = — (xf -f x22) — |m2 sgn
l+\ IT
Strong practical stability is a general result for systems of this type.39
522 Introduction to Stability Theory and Lyapunov's Second Method
Fig. 7.13-1
Y(s)
Fig. 7.13-2
Sec. 7.13 Relay Controllers for Linear Dynamics 523
Example 7.13-2. A relay controller is to be designed for the dynamics of Fig. 7.13-3a.
Y(s) 1 _ 1 1
State equations are defined for the system of Fig. 7.13-36 as x = —x + m and z = m.
Note that the first equation also applies to Fig. 7.13-3(7. z, defined in Fig. 7.13-36, is
x + y of Fig. 7.13-3(7. Assuming P = 2, then Q = 1. Since 60 = 1 and b is a scalar
Fig. 7.13-3
the system has a strong practical stability which can be made as strong as desired by
increasing M. This capability, of course, is bounded by the saturation tendencies of the
linear dynamics.
524 Introduction to Stability Theory and Lyapunov's Second Method
Lyapunov’s second method has been used in preceding sections for the
determination of stability and for the design of relay systems. Other than
for linear systems and the nonlinear case considered in Problem 7.21, no
method has been presented for the determination of Lyapunov functions.
Since various theorems guarantee the existence of Lyapunov functions for
the stable and asymptotically stable (locally or in the large) cases if f(x)
satisfies the Lipschitz condition! (locally or ih the large), it is important to
consider possible methods of determining such functions.54 At least three
methods are presented in the literature.55-60 The variable gradient method
is considered here.
Given the existence of a Lyapunov function, its gradient must exist.
From grad V, V(x) and W may be determined, since, for x = f(x), Eq.
7.8-3 indicates
The upper limit of integration does not imply that V(x) is a vector. It
indicates that the integral is a line integral to an arbitrary point x =
(xl9 x2, . . . , xn) in state space. For the scalar function L(x) to be obtained
uniquely from the line integral of the vector grad V, V(x) must be inde¬
pendent of the path of integration. The necessary and sufficient con¬
ditions for this arej
(grad V\
(grad V)2
grad V =
(grad V)n
t The Lipschitz condition implies continuity of f(x) in x.
f See any standard text on vector calculus.
Sec. 7.14 Determination of Lyapunov Functions 525
Xl
a„i(x) *n2(x) ‘ ‘ *
"allc + &iiv(xi) a12cT a12l/(a:l> X2> • • • Xn—l) • Q^lnc “1” •*'2> • • • ^n—l)
With [a] as defined above, the steps in the variable gradient method of
determining a Lyapunov function are:
grad,/
V =
,
x0
Then, from Eq. 7.14-4, V is given by
*2
V(x) = g(ux) dux + u92 du
<-<“2
For g(u) as defined in Example 7.8-7, V(x) is positive definite. The Lyapunov function
determined here is precisely the one used in Example 7.8-7.
Example 7.14-2. Determine a Lyapunov function for the system of Example 7.8-8.
The system to be considered is third order and is defined by xx = x2, x2 = x3, and
x3 = —h(xx) — g(x2) — ax3. For a third order case, grad V is
This can be satisfied simply by a12c = a21c = a13u = a23u = 0, ocm = dh{x^\dxx, and
oc2iv = h(x^)lxx. Similarly, for i = l,y = 3, and the a’s above, Eq. 7.14-3 yields
This can be satisfied simply by a13c = a31c = a31„ = a32v = 0. Similarly, for i = 2,
j = 3, and the a’s above, Eq. 7.14-3 yields a23c = a32c. This can be satisfied simply by
a23c = a32c = a. For these a’s, W is
dh(xx)
W = [allc + an^J]®!®, + —— x22
dx1
It is now apparent that some of these choices for the a’s were determined while
observing W.
Now W can be put into a rather simple form by choosing allc = 0, a11(l(®i) =
ah(x1)/x1, a22c = a2, and cc22v(x2) = g(x2)/x2. Then W becomes
g(x2) _ dh(xJ'
W = — a Xo
x« dx1
The corresponding grad V is
dh(xx)
ah(x J + Xg
dxx ~2
The V(x) determined here is precisely the one utilized in Example 7.8-8. The conditions
for asymptotic stability are given there.
dv,dv
ww= —- + — /i(x) + — /2(x) +.
dt dx±
s^_dv
dx2
\ , dV
+ —A(x)
do: n
.
dV
W= — + (grad V, f)
dt
r 0
n
a
_ y(t) a_
Fig. 7.15-1
Sec. 7.16 Eventual Stability 529
Then
2a
V(x, t) = P + Xi2 + 2xxx2 + - X
M0_ P
Let Va(x) = eyc-L2 + 2a:1x2 -f- e2x22. Va(x) is positive definite, and certainly V(x,t) >
Vaix) if
n 2a 2
P + -s-77:
Py(t) > ei > 1 and ^
p > *2 > 1
Now, considering W,
W = <x, Q(t)x) + (x, Q(t)x) + <x, Q(r)x>
<x-y(0
1 + 0
y( t) Py(t)_
P =
2
0
a
1 «y(0
1 + > 0
7(0 M0_
Sufficient conditions for satisfaction of these inequalities, and that some Vb(x) dominate
V(x, t), and hence asymptotic stability in the large, are a > y(t) > 0, 2 > ft > 0, and
7(0 > 0.
The origin of x = f(x, t) is eventually stable if, given e > 0, there exists
d and T such that ||x0|| < (5 implies \\x(t; x0, f0)|| < e for all t > t0^ T.%1
Lyapunov’s second method can be extended to study eventual stability.
LaSalle and Rath contribute several theorems to this extension. One of
these, and their example relating to adaptive control, are presented here.
\ht(t)\ dt <oo, i = 1, 2
Jo
If in addition,
4. For some K and some 0 < a < 1, \q(x, z)| < KV~a(x, z), then all
solutions z(t;z0,t0) are bounded and all x(t;x0,t0) approach 0 as t
approaches infinity.
would solve the problem, i.e., the steady-state solution would be k. Since 0(0 and 0O
are unknown, however, replace 0(0 in the brackets by an adjustable feedback gain
zx(t). Thus
xx = —0(0^1 + (zi — d)xx + ak, a > 0
Sec. 7.17 Discrete Systems 531
Let V(xx, zx) = %[(xx — k)2 + (zi — /50)2]. Then W = dV/dt is given by
IV = -a(xx - ky - 0 - p0)xi(xi ~ k)
Denoting Va(x) = a(xx - k)\ hx(t) = V2 \k\ • |P(t) - /S0|, A2(/) = 2 |^(0 - j50|, and
<7(x, z) = V F(xl5 2X), then
The equilibrium is given by the solutions of f(0, kT) = 0 for all k, and
f(x, kT) is assumed to be real and continuous with respect to x for fixed k.
The fundamental stability definitions, given in preceding sections for
continuous systems, are applicable to discrete systems. A Lyapunov
function used for the investigation of the stability of an equilibrium of a
discrete system need only be defined for the discrete instants of time kT.
Similarly, the definiteness requirements need hold only for these discrete
instants. Thus, instead of the total derivative of the Lyapunov function,
which is of interest for continuous systems, one considers the total
differencef
f W can be divided by T to determine the rate of increase of V; however, this does not
change the definiteness of the function.
532 Introduction to Stability Theory and Lyapunov’’s Second Method
f$(k) and y(Jc) are yet to be determined functions of the state variables of
the linear dynamics.
Expanding G(s) in partial fractions, as in Sections 5.5 and 7.10, x =
Ax + mb, where A is the diagonal matrix with elements A1? A2, . . . , Xn.
Linear
dynamics
Fig. 7.18-1
Sec. 7 AS Application to Pulse Width Modulated Systems 533
t0 = kT t = kT + <5 (k)
|<x, e2TAb>|
d(k) =
b, c2TAb>
534 Introduction to Stability Theory and Lyapunov's Second Method
(a)
(b)
Fig. 7.18-2
For the state variables defined in Fig. 7.18-26, <x, e2T^b) — e_2r(x1 + e 2Tx2) and
(b, e27^b> = e~2r(l + e~2T). Thus the sign of the modulator output is
sgn (1 = —sgn [xx + e_22,x2]
and the pulse width is
+ c 2r^2|
3(k) = T sat
7(1 + e~2T)j
Note that neither xx nor xx exists as such in Fig. 7.18-2#. However, by the methods of
Chapter 5, xx — 2z1 + z2 and x2 = — (zx + z2). Then, if zl and z2 can be measured by
appropriate sensors, the proper switching signals for the pulse width modulator can be
obtained from a linear combination of these measurements.
References 535
REFERENCES
26. J. LaSalle and S. Lefschetz, Stability by Lyapunov's Direct Method with Applications,
Academic Press, New York, 1961, p. 37-38.
27. Struble, op. cit., p. 161.
28. W. Hahn, Theory and Application of Liapunov's Direct Method, Prentice-Hall, Inc.,
Englewood Cliffs, N.J., 1963, pp. 14-15.
29. Ibid., p. 15.
30. Struble, op. cit., p. 164.
31. LaSalle and Lefschetz, loc. cit.
32. Hahn, op. cit., pp. 47-48.
33. E. A. Barbasin, “Stability of the Solution of a Certain Nonlinear Third-Order
Equation” (in Russian), Prikl. Mat. Mekh., Vol. 16, 1952, pp. 629-632.
34. W. J. Cunningham, An Introduction to Lyapunov's Second Method (AIEE Work
Session in Lypanuov’s Second Method, edited by L. F. Kazda,) Sept. 1960, p. 30.
35. Hahn, op. cit., p. 15.
36. LaSalle and Lefschetz, op. cit., p. 67.
37. E. A. Barbasin and N. N. Krasovskii, “Concerning the Stability of Motion as a
Whole,” Dokl. Akad. Nauk SSSR, Vol. 36, No. 3, 1953.
38. E. I. Ergin, V. D. Norum, and T. G. Windeknecht, “Techniques for Analysis of
Nonlinear Attitude Control Systems for Space Vehicles,” Aeronautical Systems
Division, Dir./Aeromechanics, Flight Control Lab, Wright-Patterson AFB, Ohio,
Rept. No. ASD-TDR-62-208, Vol. II, June 1962.
39. J. P. LaSalle, “Stability and Control,” J. SIAM, Ser. A on Control, Vol. 1, No. 1,
1962, pp. 3-15.
40. A. I. Lur’e, Some Nonlinear Problems in the Theory of Automatic Control, Gos-
tekhizdat (in Russian), 1951; English translation, Her Majesty’s Stationery Office,
London, 1957.
41. A. M. Letov, Stability in Nonlinear Control Systems, Princeton University Press,
Princeton, N.J., 1961.
42. LaSalle and Lefschetz, op. cit., pp. 75-105.
43. J. P. LaSalle, “Complete Stability of a Nonlinear Control System,” Proc. Nat.
Acad. Sci., U.S., Vol. 48, No. 4, April 1962, pp. 600-603.
44. LaSalle and Lefschetz, op. cit., p. 85.
45. J. E. Gibson, Nonlinear Automatic Control, McGraw-Hill Book Co., New York,
1963, pp. 324-326.
46. G. Franklin and B. Gragg, “Discussion of Stability Analysis of Nonlinear Control
Systems by the Second Method of Liapunov,” Trans. IRE, AC-1, October 1962, pp.
129-130. -
47. Gibson, op. cit., pp. 328-334.
48. M. A. Aizerman, Lectures on Theory of Automatic Regulation (in Russian), Moscow,
1958.
49. G. N. Dubosin, “On the Problem of Stability of a Motion under Constantly Acting
Perturbations,” Trudy gos. astron. Inst., Sternberg, 1940.
50. I. I. Vorovich, “On the Stability of Motion with Random Disturbances,” Izv. Akad.
Nauk. SSSR, Ser. mat., 1956.
51. Hahn, op. cit., p. 107.
52. LaSalle, loc. cit.
53. LaSalle and Lefschetz, op. cit., pp. 121-126.
54. Hahn, op. cit., pp. 68-82.
55. D. R. Ingwerson, “A Modified Liapunov Method for Nonlinear Stability Analysis,”
Trans. IRE, AC-6, May 1961.
Problems 537
Problems
7.1 Show that the trajectories for the linear, second order system of Fig. 7.3-1
with the outer feedback loop gain equal to zero (co2 = 0, £a> 0) are
given by x2 + 2icox1 = c.
7.2 Show that the trajectories for the linear, second order system of Fig. 7.3-1
with ||| = 1 are given by
x2 + t;0)X1 = Ce^(ox1Ux2+^co*i)} ! = +1 or -1
7.3 Sketch the phase plane trajectories for each of the cases considered in
Fig. 7.3-8 for the system of Fig. 7.3-1 where v{t) now is not zero but is
C/_i(0, a unit step function.
7.4 Derive the phase plane trajectories of Section 7.3 from the mode inter¬
pretation viewpoint of Chapter 5.
7.5 (a) Determine the location and nature of any singular points in the example
of Section 7.4 for the cases: a = 0; a > b > 0.
(b) Sketch the phase plane portraits for each of the preceding cases.
Indicate any separatrices.
(c) In which of the cases is the system on oscillator or a flip-flop?
(.d) On your sketches of part (b), indicate the value of x required to make
the circuit oscillate or flip, as the case may be.
7.6 A one-farad capacitor and a one-henry inductor are connected in series
with a tunnel diode negative resistance device characterized by i —
—v + v3. Let xx be the capacitor voltage and x2 be the current as shown
in Fig. P7.6.
538 Introduction to Stability Theory and Lyapunov's Second Method
vc ~ ^-'1
A
Negative
resistance
Fig. P7.6
(a) Determine the location and nature of any singular points, and sketch
the xxx2 phase plane portrait. Are x1 and x2 a set of state variables for
this system? '
(b) Repeat part («) in a w^-plane.
7.7 (a) Determine the location and nature of any singular points, and sketch
the phase plane portrait for the frictionless pendulum described by
6 + sin 6 = 0. Indicate any separatrices.
(b) Repeat part (a) for 6 + 0.16 + sin 6 =0.
7.8 Determine the location and nature of any singular points, and sketch the
phase plane portrait for the nonlinear system described by
y - (0.1 - ^y2)y + y + y2 = 0
/v m
+1
-1
+]
(b)
Fig. P7.9
Problems 539
diagram for the system is shown in Fig. PI.9a. The relay characteristic is
given in Fig. 1.9b.
(a) Assuming v(t) = 0, sketch the phase plane trajectories.
(b) If the motion is periodic, what is the period as a function of the maxi¬
mum velocity during the oscillation ?
(c) Is a limit cycle present?
7.10 The relay system of Problem 7.9 is modified by the addition of nonlinear
feedback. The block diagram for the modified system is given in Fig.
P7.10a. The nonlinear element has the characteristic shown in Fig. P7.106.
v(t) = 0
?o
e(t) m(t) y(t)
-A
Relay
tQ
-A
/ J
Nonlinear
q(t) element P(0
(a)
Q
A
-^p
■H
(b)
Fig. P7.10
(a) Repeat the questions of Problem 7.9 for the modified system.
(b) What is the effect of the nonlinear feedback on stability?
(c) Discuss the investigation of the stability of this system by the describing
function method.
7.11 The system of Fig. Pl.Wa has a nonlinear error detection means and
deadband in the load. The deadband characteristic is shown in Fig. P7.1 lb.
(a) Determine the location and nature of any singular points in an x±x2-
plane between the limits 1^1 < 2tt + a.
(b) Sketch the phase plane trajectories for this system in the same range,
and discuss system stability.
(c) Discuss the possibility and difficulty of analyzing this system by the
describing function method.
540 Introduction to Stability Theory and Lyapunov's Second Method
Deadband
(b)
Fig. P7.ll
7.12 For the system with coulomb friction illustrated in Fig. P7.12:
(a) Determine the differential equation characterizing the system.
(b) Sketch the phase portrait, indicating any singular region.
(c) Does a limit cycle exist?
7.13 (ci) Sketch the trajectories for the system of Fig. 7.5-5 in an ad-plane.
(b) Derive the successor function corresponding to part (a), and discuss
system stability.
7.14 In the system of Fig. P7.14, m(t) can be +1 or —1. The value of m{t) is
controlled by the switching logic, so that the transient error and error
derivative to a step input (i.e., v(t) = f7_i(0) are reduced to zero in the
shortest possible time. This corresponds to the least number of changes
of sign of m(t) and is the case of optimum switching.
Problems 541
Fig. P7.14
V(x) =
where
d2x, dx,
+ h(x'] Tt + s(Xl) = 0
(b) Using the K-function
x 2
V(x) = — + J *i
g(u) du
Vehicle dynamics
7.20 The RLC network of Fig. P7.20 consists of a passive resistance network,
and capacitors and inductors, all of which may be nonlinear. Choose as
the state variables the charge on each of the capacitors and the flux linkage
of each of the inductors. Let the F-function be the stored energy, and
Problems 543
em +1
A
£m + 2
/V
e*
determine the conditions for asymptotic stability in the large. How would
hysteresis affect your solution?
7.21 (a) Show that
f(x) = J(sx)x ds
(Is2 +45 + 1)
(i) ^ = - 853 - 115 -3
1
(ii) G(s) =
(5 + a)(5 + P)
(b) Derive the equation corresponding to Eq. 7.10-6, but which applies if
some of the A’s are complex with negative real parts. Use this result for
the cases
(53 + 652 + 125 + 8)
(i) G(s) ^ _|_ 1)^2 -j. 2s + 2)
7.23 Using a method analogous to that of Section 7.10, let P = uuT, where u
is a vector with components wl5 u2, ... , un, develop conditions sufficient
for stability in the large of the system of Fig. 7.10-lc. Assume that b0 = 0,
and that the controlled elements are second order with both A’s negative
real. Can this P be used to prove asymptotic stability in the large? Why?
7.24 Can an equilibrium be stable without being totally stable? Think of an
undamped pendulum. Can an equilibrium be totally stable without being
stable?
7.25 (a) Assuming v(t) = 0, for what values of K is the equilibrium at x = 0
of the system of Fig. P7.25 stable?
(<b) Assuming v(t) ^ 0, but is bounded with a maximum absolute value
of vm, determine values of K for which x ultimately lies within a finite
value of x =0. How does this finite value vary with vm and K?
Fig. P7.25
7.26 Discuss the effect of H(xl) on the speed of response of the system of
Problem 7.18.
7.27 Discuss the effect of T on the speed of response of the system of Problem
7.19.
7.28 Use the procedure of Section 7.13 to design a relay controller for the
system
a > 0, b > 0
7.29 Use the procedure of Section 7.13 to design a relay controller for the
system
0 1 0 "0 0 0“
A = 0 0 1 , B = 0 0 0
—c -b —a _0 0 1
S + cc
G(s) =
^(5 + 1)
Problems 545
What are the conditions on a and k( ) for which the system is asymptot¬
ically stable in the large?
7.32 (a) Use the variable gradient method to determine a Lyapunov function
for the system of Fig. 7.10-1/) where
s + 1
G(s) =
s(s + /3)
(/)) What are the conditions on /3 and k( ) for which the system is asymp¬
totically stable in the large?
(c) Resolve your answer to part (b) with Aizerman’s conjecture (see
Example 7.10-1).
7.33 Using the variable gradient method, determine F(x) and W for the system
of Example 7.4-1. Assume
In what regions of the ^c^-plane are the V(x) curves closed? What can
you say about the behavior of the system?
7.34 Investigate the stability of the system of Example 7.15-1, when a and /?
are also functions of time.
7.35 Develop the method analogous to Section 7.13 for the case in which x and
m are discrete signals.
7.36 Investigate the application of Lyapunov’s second method to the first
canonic form of Lur’e, when o, x, and 2 are discrete signals.
7.37 The linear dynamics characterized by x = Ax + Bm are such that
| AI — A| =0 has one root at A =0, negative real roots, and some complex
roots with negative real parts. The roots are distinct. These linear
dynamics are to be controlled by a controller such that the various com¬
ponents of m are given by
for (k -1- 1)T > t > kT, i.e., can change sign at the discrete time
instants kT, (k + \)T, . . . . Between these discrete instants, /?7Z- is a con¬
stant equal to +MZ or -Mf. Can the az be determined so that the con¬
trolled system is asymptotically stable in the large?
8
Introduction to Optimization Theory
8.1 INTRODUCTION
the extent to which these limitations affect a given design problem, are
indicated.
3. Inconsistent sets of performance specifications are revealed.
4. Prediction is naturally included in the procedure, because the design
index evaluates performance over the future interval of control.
5. The resulting control system is adaptive, if the design index is refor¬
mulated and the controller parameters recomputed on-line.
6. Time-varying processes do not cause any added difficulty, assuming
that a computer is used to determine the optimum.
7. Nonlinear processes can be treated directly, however, at the expense
of increased computational complexity.
^ Controlled
Compensation -- -^
elements
/V
(a)
Control
signals
Controlled
elements
(b)
Fig. 8.1-1
I f(t)g[e(t)] dt
error index
(8.2-1)
where yf and mf are the desired output and control effort, respectively;
ly. and lm. are related to limits on yt and mrespectively; <f>u{t), £u{t),
ipuit) and are time-dependent weighting factors; and yi and fii are
integers. The performance index considers the system behavior during the
future time interval t0 < t < tf, where tf may be a constant, a variable, or
infinity.
The weighting factors permit the various terms of the performance
index to be emphasized or weighted in time, depending upon the relative
importance of these terms. The terms raised to the powers yt and pLi are
penalty functions, which tend to maintain output and control signals
within prescribed limits. This is accomplished by heavily weighting these
signals, if they exceed their limits.
A unique set of weighting factors and penalty functions to satisfy
prescribed design specifications generally does not exist. Furthermore, the
selection of these quantities is unfortunately not a straightforward matter.
However, the lack of uniqueness of the weighting factors and penalty
functions does introduce a flexibility which makes their selection simpler.
From an engineering viewpoint, an efficient procedure for selecting weight¬
ing factors and penalty functions is needed. As discussed in later sections,
Ellert has partially answered this need.
The performance indices above, and many of the specialized indices
found in the literature, can be put in the form
For example, in the flight of a vehicle from one point to another with least
fuel consumption, minimization of Eq. 8.2-3 is desirable, if q(y, m, t) is
chosen as the fuel consumption per unit time. In chemical process control,
one might seek a maximum of Eq. 8.2-3. In the latter case, however,
q[y, m, t] typically would represent the instantaneous yield of the process.
As a final illustration, minimization of the time required for a system to go
from one state to another can be accomplished by minimizing Eq. 8.2-3,
with q[y, m, t] chosen to be a constant. In such a case, constraints would
exist on the maximum velocities and accelerations which can be tolerated.
Many more examples of optimization problems could be listed. How¬
ever, the important aspect of this discussion is that, even though these
problems are different, they are all closely related mathematically by the
objective of finding a maximum or a minimum of Eq. 8.2-3. Problems
of this type can be solved by Pontryagin’s method, or by the dynamic
programming techniques of Bellman.
x = f(x, m, t)
(8.3-1)
y = g(x, 0
The elements of f are assumed to be continuous with respect to the ele¬
ments of x and m, and continuously differentiable with respect to the ele¬
ments of x. The controlled elements are assumed to be observable and
controllable, i.e., all state variables are measureable, and it is possible to
excite every state of the controlled elements. The presentation here is
further limited to the special case for which there are no restrictions on the
amplitudes of the control signals or state variables. A more general pres¬
entation is given by Pontryagin et al.17
Before considering minimization (maximization) of a functional, as
Eq. 8.2-3, it is worthwhile to consider the more familiar case of minimiza¬
tion (maximization) of a function. All engineers have encountered prob¬
lems of trying to minimize (maximize) a function of a finite number of
independent variables, say 0(x). Points at which all the first partial deriva¬
tives of the function are zero are known as stationary points. If the function
is a minimum (maximum) at a stationary point, then that point is called an
extremum.
552 Introduction to Optimization Theory
If the variables of the function are not independent but are subject to
equality constraints, e.g., w(x) = 0, necessary conditions for an extremum
can be determined by Lagrange’s method of multipliers. This method con¬
sists in introducing as many new parameters (Lagrange multipliers)
p^p2, • • • (which may be regarded as the components of a vector p) as
there are constraint equations, forming the function 6C = 0(x) + (p, w)
and determining necessary conditions for an extremum from
ddc
Wi(x)
30c
dpi dpi
Lagrange’s method avoids having to solve the constraint equations for the
cc’s and substituting the results into 0(x). This is accomplished by intro¬
ducing the above additional restrictions.
The calculus of variations is also concerned with the determination of
extrema.f Rather than extrema of functions, however, the object of the
calculus of variations is to determine extrema of functionals. Section 1.4
indicates that, if x has a unique value corresponding to each value of t
lying in some domain, then x(t) is said to be a function of t for that domain;
to each value of t, there corresponds a value of x. In essence, & functional
is a function of a function, rather than of a variable. For example, f[x{t)]
is a functional if, to each function x(t), there corresponds a value of /.
The performance index / of Eq. 8.2-3 is also a functional.
If the second of Eqs. 8.3-1 is substituted into Eq. 8.2-3, the result can be
written as{
I = f
J to
/0(x, m, 0 dt (8.3-2)
Then the problem of determining an extremum of Eq. 8.3-2 for the con¬
trolled elements of Eqs. 8.3-1, is one of determining the function m(t)
which makes / an extremum, subject to the n equality constraints
f(x, m, t) — x = 0. The method of Lagrange multipliers is also useful for
h = fVo + <P, f - *» dt
J to
(8.3-3)
x = x° + Axa
m = m° + Bma (8.3-4)
p = p° + Tpa
where A, B, and T are diagonal matrices with elements oq, & and yi9
respectively, af, /?*, and yi are parameters which adjust the amount of
perturbation that the quantities xf, mf, and pf introduce into xi9 mi and
pi9 respectively. It is assumed that these perturbations are unrestricted.
From the first of Eqs. 8.3-4, it is apparent that
x = x° + Axa (8.3-5)
If Eqs. 8.3-4 and 8.3-5 are substituted into Eq. 8.3-3, Ic has its optimum
value If for a = A1 = 0, p = B1 = 0, y = n = 0, since x, m, and p
then have their optimum values x°, m°, and p°, respectively. Thus Eq.
8.3-3 has a stationary point at a = p = y = 0, and necessary conditions
for the optimum are
grad„ Ic |a—p—y=o = 0
gradp Ic |a=p—y=0 = 0 (8.3-6)
grady 1C |o=p=Y=0 = 0
Application of Eq. 8.3-6 to Eq. 8.3-3, after substitution of Eqs. 8.3-4 and
8.3-5, yields
where Xa, Ma, and Pa are diagonal matrices whose elements are the ele¬
ments of xa, ma, and pa, respectively, and Hc° is the optimum value of the
554 Introduction to Optimization Theory
integrand ofEq. 8.3-3, i.e., Hc° = f0° -f (p°, f° — x°). Integration by parts
of the second term in the first of Eqs. 8.3-7 allows that equation to be
written as
t=tf
subject to the boundary conditions x°(/‘0) = x(f0) and p°(^) = 0.| For
specified terminal conditions on x°, the latter boundary condition is
f The latter boundary conditions are a special case of the so-called transversality condition.
Sec. 8.3 Necessary Conditions for an Extremum 555
xi
into Eq. 8.3-2. xd is the desired state behavior, and £2 and Z are symmetric
matrices which are possibly time-varying.| The dimensions of £2 are less
than (n x n), unless all components of (xd — x) are included in f0% The
objective is to determine x°, m° and the dependence of m° on x° and xd.
From Eq. 8.3-10,
For the case under consideration, the two-point boundary value problem
can be converted into two one-point boundary value problems. Equation
8.4-3 consists of a set of interrelated linear differential equations for x° and
t The Euler equations together with a positive semidefinite £2 and a positive definite Z
constitute necessary and sufficient conditions for a minimum of the performance index,
for the class of problems considered here. Furthermore, the corresponding linear
optimum control system is stable (asymptotically stable if £2 is positive definite).25
X A similar statement holds with respect to Z in terms of the dimensions of m.
558 Introduction to Optimization Theory
p° = Kx° - y° (8.4-4)
Then substituting for x° from the first of Eqs. 8.4-3 and using Eq. 8.4-4
results in
Since this expression must be valid for all possible x, the conditions are
The first of Eqs. 8.4-5 is a set of first order nonlinear differential equations
of the Riccati type.20 The second of Eqs. 8.4-5 is a set of linear, time-vary¬
ing, first order differential equations.! In the case of unspecified terminal
conditions on x°, p°(^) = 0. Thus the boundary conditions on K and v°
for this case are that each of the elements of K and v° is zero at t — tf, as
indicated by Eq. 8.4-4.
Once K and v° are determined, the control law for the optimum system,
is given by substituting Eq. 8.4-4 into Eq. 8.4-2 to obtain
Thus, for this case, the control law is linear, and the controller feedback
gains K are independent of the state of the controlled elements. Further¬
more, since the control law is independent of the initial conditions of the
state variables, the system configuration as defined by Eq. 8.4-6 is optimum
for all initial conditions. Merriam, who first noted this property, refers to
this as the optimum configuration.14 Figure 8.4-1 illustrates this configura¬
tion for the general linear case.
Once m° is determined, the response of the optimum system can be
obtained from
x° = (A - BZ_1BtK)x° + BZ_1Brv° (8.4-7)
which results from substituting Eq. 8.4-6 into Eq. 8.4-1. Thus the two-
point boundary value problem has been converted into two one-point
f These equations are adjoint to the equations of the closed-loop (controlled) system.
Sec. 8.4 Linear Optimization Problems 559
Controlled elements
Fig. 8.4-1
boundary value problems. These are the solution of Eq. 8.4-5 backward in
time from t = tf to t — t0, and subsequently solving Eq. 8.4-7 forward in
time from t — t0 to t — tf.
In the nonlinear case, where the controlled elements are nonlinear and/or
the performance index is nonquadratic, it is not possible to convert the
two point boundary value problem in the above manner. Also, the opti¬
mum control law is not linear. These aspects then generally demand com¬
puter solution of the equations defining the optimum system, as is
considered in later sections.
Example 8.4-1. Determine the optimum controller according to the performance index
r*f
I= i [x1co11x1 + d dt
J to
for the first order controlled elements described by xx(t) = axlxx(t) + blxmx{t). The
system is assumed to be a regulator, so that xxd — 0. xx{tf) is unspecified.
For xd = 0, Eq. 8.4-5 indicates v° = 0. Therefore mx°(t) = — l\\Xbxxkxx(t)xp(t), where
kxx{t) is given by the solution to
from Eqs. 8.4-6 and 8.4-5. In order to determine kxx{t), let r = tf — t and kxx(tf — r) =
kxx°(r). Then
+ 0Jn
560 Introduction to Optimization Theory
Thus
£ii(liCi*PT "t“ @T)
k\\ °(T)
bix\c^T + c2e~^T)
Since ^u°(0) = 0, c2 = — Then
con sinh (It
*11°(T) =
(3 cosh (It — alx sinh fir
Therefore
con sinh P(tf — t)
k „(*) =
P cosh P(tf — t) — <2U sinh P(tf — t)
and m°(0 is given by
COll^H sinh P(tf -r 0
md(t) = —
*i°(0
£n _P cosh P(tf — t) — axx sinh P(tf — t)_
If tf is a constant, the terminal time of the performance index becomes nearer as real
time advances, assuming t < tf. In this so-called shrinking interval problem, the optimum
system is time-varying. If tf is a fixed time Tin the future relative to real time, i.e., tf —
t + T, the terminal time of the performance index slides ahead in time as real time
advances. This is called a sliding interval problem, and, if xd, SI, Z, and the linear
controlled elements are time-invariant, the resultant system is stationary. A special case
of these is given by infinite tf. This is the infinite interval problem, f If xd, 52, Z, and the
controlled elements are time-invariant, the resultant system designed according to an
infinite interval performance criterion is stationary. In this example, mP(t) becomes
COji^n
mx\t) V(0
Ui(P ~ On)
corresponding to a stationary system.
For the case in which the controlled elements consist of an integrator without feed¬
back, axl = 0. Then, for the infinite interval case, mx0(t) is
V%
m-P(t) I *i°(0
As (OnlCii is increased, so that the performance index emphasizes the system error
relative to the cost of reducing it, the loop gain increases. Also, the speed of response as
indicated by
\A
®1°(0 = xiVo) exp (t - to)
increases. This agrees with one’s intuition based on conventional feedback control
theory.
-1
G
•e*
1—
.-*r(-s)[p0(0) +£lXd(s)]_
i
-P°(s).
<
where 4>(s) = (si - A)-1 and -&T(-s) = (si + A2)-1, and X°(s), P°(s)
and Xd(s) are vectors. 3>(s) and — <!>(—s) are the Laplace transforms of the
state transmission matrices for Eq. 8.4-1 and its adjoint, respectively.
Since
-1
I a12~ (I - a12a21) 1 -(I - a12a21) la12
Example 8.4-2. Repeat Example 8.4-1, using the Laplace transform method.
For this case,
1
<*>0) = [si - A]-1 =
S - an
and Xd(y) = 0. Then
-i
(Dub ii 2 1 c°iixi0(Q)
iY(s) = - 1 - pm
£u0 - fluX-s + «n)_ s + an a ii
which is the result previously obtained by the more direct method. This result is also
illustrated by Fig. 8.4-2.
Example 8.4-3. Determine the optimum system for the controlled elements of Fig.
8.4-3. The performance index is described by
'(On O' 0 0 ■
£1 = z =
0 0 o £22
m2 = *2 X2 = *1 xi
/ -> I
Fig. 8.4-3
Sec. 8.4 Linear Optimization Problems 563
Since Z is singular, the first of Eqs. 8.4-8 cannot be used directly. The factor Z-1 in
Eqs. 8.4-8 is due to solving
Zm° + Brp0 = 0 (8.4-9)
for m° and substituting the result into Eq. 8.4-1. In this case, Z is as given above anc
"0 0 “
B =
0 1
Thus the only information contained in Eq. 8.4-9 is m2° = —(p2°fC22). But, by the
problem definition, mi° = 0. Then Eq. 8.4-9 is unchanged if Z is replaced by
1 0 "
0 £22
Therefore Eqs. 8.4-8 can be used if Z-1 is replaced by Z1 1.f Then since
’5-1 s~2'
3>Cs) =
0 j -1
P°(.y) can be determined to be
0 W11
— 5“ —5
0
<0
1
5 1
.S2 —53
P°(5) =
54 + (wn/^22)
co M CO 11
G(s) =s2 + (2)W I -i-11 s +
22, £ 22,
This shows that P°(j) has two right-half-plane poles and two left half-plane poles,
symmetrically located with respect to the origin. In order to have p°(oo) = 0, the
residue in the right half-plane poles must be zero. A partial fraction expansion of either
Pi\s) or P2°(s) reveals that the requirements on p°(0) for zero residue in each of the
right half-plane poles of P0(.s) are
(4cOn3^22)/4 (W11^22)/2
P°(0) = x°(0)
(con^22)/2 (4con£223)/4
Since
s3 J2
1 —5
COli x°(0) + £2r4 P°(0)
-S3 5 — s2
£22 J
X°(j)
^4 “b (^11/^22)
t Note that Zt is positive definite and hence the resulting linear optimum system is
asymptotically stable, since the other requirements previously given for this are also
satisfied.
An obvious alternative to this procedure is to rederive Eq. 8.4-8, but for the case in
which B is a vector.
564 Introduction to Optimization Theory
7T
(2)^ sin ( at + - a-1 sin at
<\>(t) = €~at
7T
■2a sin cct ■(2)^ sin ( at — -
This is the control law for the controlled elements of Fig. 8.4-3. As expected, it is a
linear function of the state variables.
As a is increased, the performance index emphasizes the error relative to the cost of
reducing it. From <\>(t) or G(s), it can be seen that the effect is to increase the speed of
response and the natural frequency of the system. The damping ratio, however, remains
constant at 0.707. Increased damping would have been obtained if co22 > 0 had been
chosen in the performance index.
4*ll(b A) *1*12(A A)
4*(b *o) = (8.4-10)
_4*2l(A fi)) Ct>22(b fi>)_
Since 4>(f0, t0) = I,
4*n(fi)> fi)) = 4*22( A? A) = I
4*12(20? A) = 4*2i(fi)> A) = [o]
In terms of <J>(7, t0), the solution of Eq. 8.4-3 is
Then, from Eq. 8.4-2, the definition of b2(//51), and the fact that
The resulting response x°(t) can be found from Eqs. 8.4-11 and 8.4-13.
For specified terminal conditions on x°(t), i.e., x°(tf) = x(tf), Eq. 8.4-11
gives
x°(tf) = 4>11(t/, t0)x°(t0) + 4>i2(7/, t0)p°(t0) + bi(t/? t0)
Then
P°(70) = -4>i2_1(t/, t0)[4*n(t/, t0)x°(t0) - x°(tf) + b1(r/, t0)] (8.4-15)
Substituting t for t0, using Eq. 8.4-2, the definition of b^, t) and the fact
that
4*12 04*i2(7/, t) == 4*12(7, t)
yields for the control law, in the case of specified terminal conditions on
x°(0,
The resulting response x°(t) can be found from Eqs. 8.4-11 and 8.4-15.
566 Introduction to Optimization Theory
The reader should note that these control laws are of the form of Eq.
8.4-6, where
K = <j>22 1{tf, 0<4>2l(^/? 0
for specified x°(//). Also, the control law requires knowledge of xd in the
future interval of control, i.e., xd(r) for t < r < tf. This is a general
requirement for optimization according to this performance criterion.
Example 8.4-4. Repeat Example 8.4-3 in the time cjpmain.
Let
" 0 -1 0 0 ■
0 0 0
" A -BZ-1Br“
G = = (-l)
-SI -Ar
COn 0 0 0
_ 0 0 1 0
so that 4>(Y, /0) = eG(<-<oh Use of the Cayley-Hamilton technique gives
1_
CLx(t - t0)
8
W11C22 la3(/ t0) cc0(t - t0) ^22 la2^ ^0) £22 1<xi(t to)
4>(h t0) =
o)iiCCi(t r0) co11oc2(t /0) cc0(t - to) ft>11^22 1(X3(t ~~ to)
_co11<x2{t t0) wna3(/ io) — <*i(t - t0) oc0(t - r0) -
where
a0(O = cosh at cos cut
sinh cut cos at + cosh cut sin cut
ai(0 =
2a
sinh cut sin cut
a2(0
2a^
cosh cut sin cut — sinh cut cos at
a3(0
4od
and a is as defined in Example 8.4-3. Substitution into Eq. 8.4-14 yields, for infinite tf,
m2 °(t) = —2a[ax1°(r) + ^2°(01
Example 8.4-5. Determine the optimum system for the controlled elements character¬
ized by
1 1
x1 = — - xx + - mx
/ = \ f ixi + m i2) dt
Jt0
and xxd{t) = 0. The terminal conditions are unspecified.
Sec. 8.4 Linear Optimization Problems 567
_
>
-1
1
Then the components (fruit, t0) and (fr21(t, t0) of 4>(r, /0) must satisfy
Solving for (fruit, t0) in the second equation and substituting the result into the first
yields
This is a form of Euler’s equation, considered in Example 2.8-2. The change of variable
t = ez gives
021 'V, eZ°) - 021V, <*») - 0.l(€«, e20) = 0
where the primes denotes differentiation with respect to 2. The differential equation
for 021(ez, ez°) has the solution
/ V5 + 1 \ /l — V5
02 ii*z, eZ°) = ki(zo) exp -2 + k2iz0) exp I-2
2 / \ 2
so that
(f>2lit,t0) = C1(^V(V5 + 1)/a + C2(r0)?F-V 5)/2
Similarly,
1 - V5' t(V*-i)/« + e2(;0)^ + V 5j r(Vi+i)/2
(fruit, t0) —
Since 011(7O, f0) = 1 and 02i(/o, *o) = 0, cx(r0) and c2(/0) can be determined to be
1
_ ?(l-V5)/2
Clito) —
V5 °
1
C2it0)=--t^V 5)/2
V5
Then
, . V5 - 1\ /t0V(V5-l)/2 /V5+ l\//0\(V/5+l)/2
Similarly,
r1 /to\-(V5-l)/2 t~1 /^o\(^5+1)/2
(fruit, t0) —
568 Introduction to Optimization Theory
mx°(t) =-aq°(0
t ^22(^5 ^/)
Thus
V5
1 - (///,)
mx°(t) = —2 *1 °(0
_(V5 + 1) + (V5 - 1)(////)VI_
mAO = - V(0
V5 + 1
In this case, a time-invariant control law is obtained, even though the controlled
are time-varying.
Example 8.4-6. Determine the optimum system for the controlled elements of Example
8.4-5, if the performance index is
/ = £ j wij2 dt
Jt0
0 t -1
4>(r, t0) is the solution to 4>(7, /0) = G<J>(/, t0). The equations can be integrated by
separation of variables to yield
to 1 1
t t to
<\>(t, to) =
t
0
to
From Eq. 8.4-16,
md(t) = - *i °(0
tf - t
to(tf — t)
*i°(0 = , ., '
t\tf to)
The response does satisfy the terminal condition xx°(tf) = 0. For t > tf, the response is
not zero, however. This is to be expected, since the performance index does not consider
this part of the response.
Although the time-varying feedback gain becomes infinite at t = tf, mx°(t) is always
finite. In fact, from the expressions for mx°(t) and ^(r).
™i°(0 = ~
tf 10
S<?c. 8.5 Selection of Constant Weighting Factors 569
The examples of the previous section indicate that the control law and
system response are greatly influenced by the weighting factors £2 and Z
chosen in the performance index. Selection of these weighting factors is a
difficult task, since the relationships between the weighting factors and
the optimum system parameters or the system response are generally
very complex. However, Ellert has developed a technique for the selection
of weighting factors in the time-invariant case.16
Consider, as an example, the second order linear controlled elements
described by
a ii *12 "0 0 “
x = X + m
L*21 *22 _0 b22_
The performance index is the one of Section 8.4, with infinite tf and
a>n 0 "0 cr
£2 = Z =
0 (x>22_ _0 i_
Using the method of Eqs. 8.4-5, the optimum control law is found to be
and is defined by
Xi\s) _
(8.5-4)
F(s) s2 + z-lOJqS + co02
where
%co0 = b22 k22 cin -*r a22
a)02 = an(a22 b222k22) + ci12{b22k21 a21)
js/ x h 2
ci12o22
V(s) = -7- ^lO)
o)0
These expressions determine con and co22, once values of z1 and co0 have
been selected.
Ellert’s procedure is to choose z1 to provide the desired relative stability
of the system, assuming that none of the system variables exceed their
prescribed limits. co0 is then chosen in accordance with the system band¬
width requirements or any limits on m2(t). The relationship between m2{t)
and co0 is given by substituting Eq. 8.5-5 into Eq. 8.5-1. It is
Table 8.5-1
Maximum
System Percent
Type Standard Forms Overshoot
r(t)
r(t)
r(t)
Sec. 8.6 Penalty Functions 573
As indicated in Sections 8.3 and 8.4, the solution of Eq. 8.3-12 requires
the solution of a two-point boundary value problem. It is not possible
to solve simultaneously the x and p equations forward in time, unless
correct boundary values for p(/0) are known. A similar problem exists if
one attempts to solve the equations simultaneously backward in time.
One possible approach is to assume a set of values for p(f0), indicated by
p1^). The superscript “1” is used to denote the first choice of p(f0).
These n conditions, in conjunction with the n conditions x°(/0) = x(/0)
are sufficient to solve Eq. 8.3-12 forward in time to determine x\t) and p 1(t).
The computed values of p1(^/) or as the case may be, can then be
compared with the correct values specified by the terminal boundary
conditions. If these two sets of values are identical (this would indeed be
fortunate), pl(t) = p°(f) and x\t) = x°(t), and the problem is solved.
Generally, however, they differ. If the x and p equations are linear,
superposition can be used to directly revise p1^) so that the terminal
boundary conditions are satisfied, and hence immediately yield the
optimum solution. In the more general case, a revised choice for p(/0),
namely p2(r0), must be made and the equations solved again. This process
is repeated until the terminal boundary conditions are satisfied, to an
acceptable degree of accuracy. This technique is called boundary condition
iteration.
More direct methods for performing boundary condition iteration are
suggested by Merriam, Neustadt, Scharmack, and Speyer.25-28 For
example, a hill-climbing problem can be formulated so that the optimum
occurs when pn(tf) = p°(A). Thus the computer techniques for hill¬
climbing problems can be utilized.
There are two significant problems associated with boundary condition
iteration methods in the general case. First, as indicated by Eq. 8.4-8 for
t In this case, the first footnote of Section 8.4 still applies, if the terms “positive semi-
definite” and “positive definite” are replaced by “convex and differentiable” and
“strictly convex and differentiable,” respectively.25 Of course, these terms now apply
to the appropriate terms in the integrand f0 of Eq. 8.3-2, rather than to £2 and Z.
Sec. 8.8 Dynamic Programming 575
Fig. 8.8-1
576 Introduction to Optimization Theory
where R[x(kj), m(k), kj] denotes the portion of /* due to kj < k < kf.
Defining a policy to be any rule for making decisions which yields an
allowable sequence of decisions, Bellman’s principle of optimality can be
stated as, “An optimal policy has the property that, whatever the initial
state and initial decision are, the remaining decisions must constitute an
optimal policy with regard to the state resulting from the first decision.”
In essence, this is a statement of the intuitively obvious philosophy that,
if /* is to be minimized by a choice of m(&), the portion of /* denoted by R
must be minimized. Furthermore, this principle states that, if R is
minimized by a selection of the sequence m(k), kj < k < kf, the optimum
sequence m°(k) is a function of the states x{k0). The state of the controlled
elements at k = kj determines m°(k), k}< k < kf. This functional
dependence can be indicated by m°(/c) = gfc[x(/c;)]. Thus the minimum
value of R is a function only of the state of the controlled elements at kj9
and kj itself. Then
dl°[x°(r), t]'
/0[x°(r), m0(r), r] + = 0
dr
where t0 < r < tf, and 7° is the minimum value of the performance index.
The total time derivative can be written in terms of partial derivatives,
and, using the first of Eqs. 8.3-1, this expression is
rc+1 = y
J to
'[/‘+1 + <p2+i, r+i - x2+1>] dt
Substitution of .,, . .
ft1 =/.*+ 4/7
pi+1 = p‘ + dp'
f«+i = fi +
xm = x‘ + dx‘
yields
The first term in the integrand gives Ic\ The last term is zero by the first of
Eqs. 8.3-1, applied to the zth iteration. Thus
'if
Ic =i: + \ [Sfo + <p‘, -5f4> + <V> <5f*) - <p\ dr) - (V, <5**>] dt
•l to
(8.9-1)
If /J+1, P*+1> and V+1 are expanded in Taylor series about their values
in the preceding iteration,
where F(x*) is the Jacobian matrix of Eq. 7.4-4, for f = f* and x = xi.e.,
the zth iteration. F(nT) is the corresponding matrix with x* replaced by nT.
P(xl) and P(nT) are matrices which correspond to F(x2) and F(m*), with f
replaced by p. Substituting these expansions into Eq. 8.9-1 and neglecting
all terms in (5nT and dx1 above first order yields
The bracketed quantity in the integrand is (d/dt) (p*, dx*). Since (5x?(/0)
must be zero, and since either pl(tf) or dxl{tf) must be zero, the result of
integrating the bracketed quantity is zero. Thus
(5nT) dt (8.9-4)
It should also be indicated that the algorithm for changing the control
signal ignores the fact that the shape of the surface on which the optimum
value of the Hamiltonian function H lies changes as x and p change.
This reduces the rate of convergence and contributes to the difficulty in
determining the
The minimization of
where w4 is defined by
TzwJ = — sz — FT(ml)g (8.10-8)
582 Introduction to Optimization Theory
f The computer results presented were provided by Dr. F. J. Ellert and were obtained
using a computer program developed at the General Electric Research Laboratory
under the direction of Dr. C. W. Merriam, III.
Sec. 8.11 Design Example 583
52
(b)
Fig. 8.11-2
cutting roller speed. In order to avoid damage to the material being fed
to the shear, the speed transition must be smooth. However, the transition
should be rapid to reduce the amount of material of undesired length
produced during the transition, since such material is waste. The desired
transitional speed versus time is shown in Fig. 8.11-2#. For simplicity,
the initial speed is normalized to unity, and the transition is to zero.
The controlled elements are represented in Fig. 8.11-2b. The variable
y1 is the cutting roller speed. The roller drive motor torque is y2. Thus the
equations for the controlled elements are
x = Ax + Bm
(8.11-1)
y = Cx
where
1
1
o
<N
Fig. 8.11-3
Fig. 8.11-4
Sec. 8.11 Design Example 585
The system saturates outside these ranges. The initial condition response
of the system should be slightly underdamped, with the peak overshoot
not exceeding 5 percent. The ranges of initial conditions are
A
0.16 -
0 /.I I 1 I t
1 8 10
/
/
- 0.16
-0.24
Fig. 8.11-6
Sec. 8.11 Design Example 587
With these expressions, and those for gradx* f0\ gradm* f0\ F(xz), F(nE),
and P(xz), the equations for the computer solution can be determined.
The first forward time expression is obtained by rewriting Eq. 8.10-9 in
the form wz+1 = ew2 + v2 — k21lxi+1 — k22x^+1, where generally y* =
m? + K*x*. In this specific case, v2 = mj + k^xp -f k22ix2i. The y
equations are solved backward in time. The forward time state equations
can be written from Eq. 8.11-1 as
V+1 ,*+i
'2
The backward time expressions for p are obtained from Eq. 8.10-3 as
Fig. 8.11-8
Sec. 8.11 Design Example 589
Fig. 8.11-10
590 Introduction to Optimization Theory
curve of Fig. 8.11-7 does have an undesirable dip. This is due to the fact
that the optimum was based on the initial conditions corresponding to the
dashed curve. In design problems such as this, in which the range of
initial conditions is large, a linear controller is limited with respect to the
performance it can provide. In such cases, the desirability of a nonlinear
controller is indicated. Design methods for such controllers are still under
investigation, however.25
In Section 8.3 it was stated that m°(t) must be chosen from a set of
bounded functions. The possibility of forcing m°(t) to be bounded by use
of penalty functions is indicated in Section 8.6. Presently of somewhat
more mathematical than practical interest is the design of a system with a
hard constraint on m°(r) without employing penalty functions. For
example, if the equations describing the optimization problem, including
the zeroth coordinate introduced because of the performance index, are
of the form
which is adjoint to Eq. 8.12-2. Since Eq. 8.12-4 is linear and does not
contain x° or m°, the general form of the solutions can be easily determined.
Elowever, the boundary conditions on p° are unknown.
Since Eq. 8.3-12 also yields
“ o r "0 O'
and B =
—b —a_ _0 1_
s2 1
where Si and s2 are the characteristic values
2
a
- b
592 Introduction to Optimization Theory
and for simplicity it is assumed that (a/2)2 — b is positive. Then Eq. 8.12-2 becomes
w = Aw + S-1 Bm, where
S2 O' "0 r
A = and S-1B =
_0 Si_ _0 1_
pi — s2px
p2 = ~Sip2 (8.12-6)
and Eq. 8.12-5 indicates that (S_1B)Tp = 0 at the switching instants. Thus the
switching instants are given by
/>i(0 +/>2(0 = 0 (8.12-7)
Since Eq. 8.12-6 corresponds to Eq. 7.3-11 except for the minus signs, the pxp2 trajec¬
tories are the zxz2 trajectories of Fig. 7.3-4a, except that the direction of the arrowheads
must be reversed. Superimposing the line px + p2 = 0 on the trajectories of Fig.
7.3-4a reveals that Eq. 8.12-7 is satisfied once for thepxp2 trajectories in the second and
fourth quadrants, and not at all for the first and third quadrant trajectories. Thus the
optimum controller for this system switches once for some initial values of x, and not
at all for some other initial conditions. This is a useful piece of information, which later
helps to define the optimum controller. It is a special case of the general result that the
time optimal control for an /?th order linear system with all real characteristic values has
no more than n — 1 switching instants.34 Unfortunately, this is the only information
provided by Eq. 8.12-5. The optimum controller must be determined from this informa¬
tion and geometrical considerations involving the xxx2 trajectories.
For the case in which (a[2)2 — b and a are both positive, the xxx2 trajectories have
the form of those for the stable node illustrated in Fig. 7.3-8. They are displaced
horizontally, however, since the optimum m2 is not zero, as assumed in Fig. 7.3-8, but
This moves the singular point. For px + p2 positive, the displacement is MJb to the left.
For px + p2 negative, the displacement is M2/b to the right. These cases are illustrated
Fig. 8.12-1
Sec. 8.12 Singular Control Systems 593
in Fig. 8.12-1. Since the two trajectories labeled “switching curve” are the only tra¬
jectories which pass through x = 0, it is apparent that the optimum x(t) must reach
x = 0 by one of these two trajectories. If the initial point is on one of these two
trajectories, no switches of m° are required. The system follows the trajectory to
x — 0.
If the initial conditions do not correspond to one of the switching curves in the
segment before the particular curve intersects the origin, then the motion of the system
must be to follow one of the trajectories upon which the initial point lies. There are
two of these trajectories, corresponding to px + p2 positive or negative. If the motion is
to reach the origin in one switch, however, the motion must follow the trajectory which
intersects a switching curve. Only one of the two trajectories passing through the
initial point does this, as can be observed by superimposing Figs. 8.12-1 a and 8.12-1 b.
This dictates the proper sign of px + p2 until the motion reaches the switching curve.
At this instant, m2°{t) must switch sign, and the motion then follows the switching curve
to the origin. This is the only manner in which the origin can be reached in one switch,
corresponding to the requirement discussed above. From these considerations, the
optimum trajectories are concluded to be those of Fig. 8.12-2, determined by combining
the appropriate portions of Figs. 8.12-1 a and 8.12-16. For the upper switching curve,
and for all points to the right of the switching curves, the optimum m2{t) is m2°(t) =
— M2. For the lower switching curve, and for all points to the left of the switching
curves, m2°(t) = M2. If m2°(t) is as specified by these two equations, x(t) reaches the
origin in least time.
594 Introduction to Optimization Theory
by
(—s2x 1 + x2) = c( —S1X1 + x2)sl's2
if x1 is replaced by xx + 1 for m2°(0 = —M2 and x1 — 1 for m2°(t) = M2, and c is then
determined so that xx — x2 = 0 is a solution of the equation. The result is
for m2\t) = — M2, i.e., the upper switching curve of Fig. 8.12-2. The lower switching
curve is described by
These expressions, and measured values of xx{t) and x2(t), could be utilized by a special
purpose “computer” to generate m2°{t).
It is interesting to compare the controller determined in this example with the one
determined in Example 7.13-1, since the controlled elements are the same. Since the
switching function of Example 7.13-1 is restricted to be a linear combination of the state
variables, the switching curve is the straight line x2 = —axJi 1 4- b). This switching
line is shown in Fig. 8.12-2, for a = 3^ -f 3-1^, b = 1. It results in more than one
switch to reach the origin, and correspondingly the motion takes longer time. Typically,
the motion would “chatter” along the switching line to reach the origin.
Example 8.12-2. The controlled elements for the single-axis attitude control of a space
vehicle are approximated by the representation of Fig. 8.12-3. Determine the optimum
controller to remove initial condition errors according to a minimum of the performance
index
m2 = X2 X2 = Xl *i
-^ /
/
Torque Rate Angle
Fig. 8.12-3
Sec. 8.12 Singular Control Systems 595
X2
*2 X2
Fig. 8.12-4
-pi = 0
— p2 = Pi + m2 sgn (m2x2)
H is minimized by
(0 if \x2\ > \p2\
m2 =
\-M2 sgn(/?2) if \x2\ < \p2\
and m2 switches when \x2\ — \p2\. This problem differs from Example 8.12-1 in that a
region of state space is introduced for which m2 — 0. This is a result of the energy term
in the performance index.
Since dx2/dxl = m2/x2, the trajectories are given by
x2 — constant, m2 = 0
x2 2
+ [M2 sgn (p2)]x 1 = constant, m2 0
2
They appear as in Fig. 8.12-4. Since only two of the trajectories pass through the origin,
they must comprise an optimum switching curve, as shown in Fig. 8.12-5a. However,
596 Introduction to Optimization Theory
there must be a second switching curve, since the p2x2 trajectories are vertical lines for
m2 = 0, indicating two values ofp2 for which \p2\ = \x2\. These trajectories are illustrated
in Fig. 8.12-56, and they also show that the change in p2 between the two switches is
equal to twice the value of x2 at which the switches take place. This information,
combined with another of Pontryagin’s equations, determines the other switching curve.
Pontryagin has shown that H is a positive constant or zero along an optimum tra¬
jectory, if H does not contain t explicitly.17 Furthermore, if tf is not fixed, as it is not in
this case, H° = 0 at t = tf. If this were not true, a further minimization of I could be
obtained by increasing tf. Thus H° = 0 for all t0 < t < tf. In particular, in the region
where m2° = 0,
H0 = 0 = p1°x2s + c
where x2s is the value of x2° when m2° = 0, i.e., w'hen the switches occur. But, if
m2° = 0,
dp2° _ pi _ c
= constant
dx!° a2s x2s2
*2
Fig. 8.12-5
Sec. 8.12 Singular Control Systems 597
/.x2s
7 IT ^
<5V
c
The latter relationship results from the fact that the change in p2° between the two
switches is 2x2s. Since the previously determined switching curve is given by
o
V/
— + M2xl = 0, x2 > 0,
rH
/v> 2
o
A
- M2x1 = 0, x2 < 0,
rH
L
/y»
Xo
2
it 2x23\
o'
<M
2 \
V c /
ry> 2
/ 2x23\
o'
<0, x2 > 0
7
II
h+ c
<M
The two switching curves and typical trajectories are shown in Fig. 8.12-6.
It is interesting to note that, as c approaches infinity, the two switching curves coalesce
and the region for which m2° is zero disappears. This is the time optimal controller for
this case, since the energy term in I is negligible compared to the infinite weighting on
minimum tf. (See also Problem 7.14.) As c approaches zero, the latter switching curve
becomes the horizontal axis. The system is turned off*. This is obviously the way to
save energy, if one does not care how long it takes to reach the origin.
x2
Fig. 8.12-6
598 Introduction to Optimization Theory
REFERENCES
Problems
I =
rtf/0(x, rn, r) dr
a =
Yfis) K
Mz(s) s(Ts + 1)
/ = (
{[yi (r) - Vi t)]2 + £22"V(t)} dr
Problems 601
(a) The system is to be designed for optimum slewing operation. For this
case, is taken as a step signal of amplitude a. Assume K = 0.2,
K[ 1,22^ ~ 2 and T = I/V5. Determine m£(t) and y^it).
(b) Repeat part (a) for the case in which |ra2(OI < M, where
8 2
M =- =- = 7 t
1 + V5 2 + V5
Compare m2°(t) and y^it) with those of part (a). Discuss the effect of
decreasing K on these responses.
(c) For optimum tracking operation, y1d(t) is taken as a unit ramp.
Repeat parts (a) and (b).
8.6 Derive Eqs. 8.5-1, 8.5-2, and 8.5-3.
8.7 Derive Eqs. 8.10-6 and 8.10-7.
8.8 Repeat Example 8.12-1 for (a/2)2 — b < 0, a > 0.
8.9 Repeat Example 8.12-1 for (#/2)2 — b > 0, a < 0.
8.10 Sketch the curves of Fig. 8.12-6, taking into account the fact that for a
space vehicle
%\(t/) — 2.nk , k =0, i 1, i2, . . . , x2(tf) = 0
are all acceptable terminal conditions.
8.11 Derive the discrete form of Pontryagin’s equations.
V
\
/I
• ' <
Index
603
604 Index
Quadratic form, positive definite, 266-270, State equations, linear RLC networks, 330-
505, 520-521 332
positive semidefinite, 267-268, 505 normal form, 340-344, 426-428
table, 268 partial fraction technique, 336-340,
Quantized signal, 7 419-426
standard form, 335
Ramp function, 12
time-varying differential equations, 397—
Random signal response, 383-385
398
Rath, R. J., 529-530
State of a system, 325-329, 413
Rayleigh equation, 491
State space, 328
Realizable system, 3
State transition matrix, continuous sys¬
Reciprocal basis, 222, 242, 346, 429
tems, 356-360
Relay controllers, 520-524
discrete systems, 432-441
Residue, 131-134
periodic systems, 372-374
Residue theorem, 132-133
properties, 369, 434
Resonance, 56, 67
time-varying continuous systems, 362-
Riccati equation, 558, 561
369
Saddle point, 477-478 Stationary point, 551
Sampling property, 16, 31 Step function, 11-12
Sampling theorem, 7, 166-167 Step response, 19, 32-34
Scharmack, D. K., 574 Stored energy, 16, 67, 121-122
Schwarz inequality, 215, 310 Strong practical stability, 518-519, 521
Semidefinite forms, 267-268, 504 Summation by parts, 444
Separatrix, 488 Superposition integral, 32-33, 376
Shannon, C. E., 7, 167 table, 33
Shaping filter, 384-385, 388 Superposition principle, 3, 23
Shifting operator, 75 Superposition summation, 38-43, 442
Signal, continuous, 6 table, 42
discrete, 6 Sylvester’s theorem, 276-281, 357, 505
quantized, 7 System, anticipative, 3
Simulation diagrams, continuous systems, deterministic, 3
314-322 fixed, 4, 48
discrete systems, 407-410 linear, 3
Singular control systems, 590-598 nonanticipative, 3
Singularity, 132 nondeterministic, 3
Singularity functions, 11-13, 18-19 probabilistic, 3
Singular point, 471-488 realizable, 3
table, 480-481 time-invariant, 4
Singular solution, 471 System function, 61, 93, 104, 122-123,
Spectral function, 9-10, 154, 157 157, 176-177
Spectral representation, 399 time-varying, 149-153, 181
Speyer, J. L., 574
Stable equilibrium, 503 Taylor’s series, 129-131
Stable system, 67, 92, 146, 447-448 Telescopic series, 78, 79
Stability, 501-504 Terminal control system, 380
Stability theorem, 506-507 Time-invariant system, 4
State determined system, 328 Time optimal systems, 590-598
State equations, 328 Total stability, 518
linear continuous systems, 329-344 Trajectory, 328, 470
linear discrete systems, 413-428 Transfer function, see System function
608 Index