DR Husari Control Notes
DR Husari Control Notes
Al-Husari
1 Introduction 1
1.1 Historical review . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Control system fundamentals . . . . . . . . . . . . . . . . . . . . 2
1.2.1 Concept of a system . . . . . . . . . . . . . . . . . . . . . 2
1.2.2 Open-loop systems . . . . . . . . . . . . . . . . . . . . . . 3
1.2.3 Closed-loop systems . . . . . . . . . . . . . . . . . . . . . 4
1.2.4 The control problem . . . . . . . . . . . . . . . . . . . . . 4
1.2.5 Examples of control systems . . . . . . . . . . . . . . . . . 5
1.3 A first analysis of feedback . . . . . . . . . . . . . . . . . . . . . 8
1.3.1 Effect of feedback on tracking and disturbance . . . . . . 9
1.3.2 Effect of feedback on sensitivity . . . . . . . . . . . . . . . 9
1.3.3 Effect of feedback on stability . . . . . . . . . . . . . . . . 10
1.3.4 The cost of feedback . . . . . . . . . . . . . . . . . . . . . 10
2 System Modelling 11
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2 The Laplace transform . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2.2 Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2.3 Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2.4 More examples . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2.5 Inverse Laplace transform . . . . . . . . . . . . . . . . . . 14
2.2.6 Partial fraction expansion . . . . . . . . . . . . . . . . . . 14
2.3 Transfer Functions . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.4 Models of Electric Circuits . . . . . . . . . . . . . . . . . . . . . . 18
2.5 Models of mechanical systems . . . . . . . . . . . . . . . . . . . . 21
2.5.1 Translational motion . . . . . . . . . . . . . . . . . . . . . 22
2.5.2 Rotational motion . . . . . . . . . . . . . . . . . . . . . . 23
2.6 DC motors in control systems . . . . . . . . . . . . . . . . . . . . 25
2.7 System modelling diagrams . . . . . . . . . . . . . . . . . . . . . 27
2.7.1 The block diagram . . . . . . . . . . . . . . . . . . . . . . 27
2.7.2 Signal-flow graph . . . . . . . . . . . . . . . . . . . . . . . 34
2.7.3 Conversion from block diagram to SFG . . . . . . . . . . 43
2.7.4 Construction of block diagrams Examples . . . . . . . . . 45
iii
iv CONTENTS
4 Stability Analysis 73
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.2 Bounded-input bounded-output stability . . . . . . . . . . . . . . 74
4.2.1 Relationship between characteristic equation roots and
stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.3 Routh-Hurwitz Stability Criterion . . . . . . . . . . . . . . . . . 77
4.3.1 General properties of polynomials . . . . . . . . . . . . . 77
4.3.2 Routh-Hurwitz stability criterion . . . . . . . . . . . . . . 78
4.4 Applications in feedback design . . . . . . . . . . . . . . . . . . . 82
1
2 CHAPTER 1. INTRODUCTION
Example 1.1 In the case of the ship shown in Figure 1.2, the rudder and engines are the
control inputs, whose values can be adjusted to control certain outputs, for
example heading and forward velocity. The wind, waves and current are dis-
turbance inputs and will induce errors in the outputs (usually called controlled
variables) of position, heading and forward velocity. In addition, the distur-
bances will introduce increased ship motion (roll, pitch, etc.) which again is not
desirable.
Control systems can be divided into two categories: the open-loop and the
closed loop systems.
Figure 1.3 represent an open-loop control system. Note that the system input
does not depend on the output, i.e., the input is not a function of the output.
A very simple example of an open-loop system is that of the clothes washing
machine. Here, the control signal is the input to the washing machine and
forces the washing machine to execute the desired preassigned operations, i.e.,
water heating, water changing, etc. The output of the system is the quality
of washing, i.e., how well the clothes have been washed. It is well known that
during the operation of the washing machine, the output (i.e., whether the
clothes are well washed or not) it not taken into consideration. The washing
machine performs only a series of operations contained in the control input
without being influenced at all by the output.
The main problem with open-loop control is that the controlled variable
is sensitive to changes in disturbance inputs. So, for example if a heater is
switched on in a room, and the temperature climbs to 20◦ C, it will remain at
that value unless there is a disturbance. This could be caused by leaving a door
to the room open, for example. The internal room temperature will change. For
the room temperature to remain constant, a mechanism is required to vary the
energy output from the heater.
4 CHAPTER 1. INTRODUCTION
The controller and plant lie along the forward path, and the sensor in the
feedback path. The measured value of the plant output is compared at the
summing point with desired value. The difference, or error is fed to the controller
which generates a control signal to drive the plant until its output equals the
desired value.
• disturbance rejection
• sensitivity reduction
• closed-loop stability
• tracking improvement
• model building
• controller design
• simulation testing
• hardware testing
The driver has the task of keeping the car on track on a desired direction of
travel. The driver uses the difference between the actual and desired direction
of travel to generate a controlled adjustment of the steering wheel as shown in
Figure 1.5.
As we all know, we use our ears not only to hear others but also to hear ourselves.
Indeed, when we speak, we hear what we are saying and, if we realize that we
didn’t say something the way we had in mind to say it, we immediately correct it.
Thus, human speech operates as a closed-loop system, where the reference input
is what we have in mind to say and want to put into words, the system is the
vocal cords, and its output is our voice. The output is continuously monitored
by our ears, which feed back our voice to our brain, where comparison is made
between ourintended (desired) speech and the actual speech that our own ears
hear (measure). If the desired speech and the measured speech are the same, no
correction is necessary, and we keep on talking. If, however, an error is realized,
e.g., in a word or in a number, then we immediately make the correction by
saying the correct word or number.
The proper procedure for teaching has the structure of a closed-loop system.
Let the students be the system, the teaching material presented by the teacher
the input, and the degree of understanding of this material by the students
the system’s output. Then, teaching can be described with the block diagram
of Figure 1.8. This figure shows that the system’s output, i.e., the degree of
understanding by students of the material taught, is fed back to the input, i.e.,
to the teacher. Indeed, an experienced teacher should be able to sense (measure)
if the students understood the material taught. Subsequently, the teacher will
Taking a shower
Imagine taking a shower from a two-knob faucet as shown in Figure 1.9. You
want to set the rate of water flow and its temperature so that the shower is
effective and comfortable. You can control the hot water flow and the cold
water flow separately.
2. Decide if it is OK
y = Gu + d (1.1)
u = Ce
e = r − (GCe + d)
1
=⇒ e = (r − d) (1.3)
1 + GC
1.3. A FIRST ANALYSIS OF FEEDBACK 9
y = Gu + d
On the other hand, with feedback control the output was given in (1.4) as
GC 1
y= r+ d
1 + GC 1 + GC
Without control the map from the disturbance d to the ouput y is simply 1.
Any disturbance d will be seen directly (without reduction) on the output y.
With control, the map from disturbance d to output y is
1
1 + GC
therefore
any disturbance
d will be seen with a factor on the output y. Hopefully,
1
the factor can reduce the disturbance. It can be seen that if C 1
1 + GC
(large or high loop gain), then
1
≈0
1 + GC
which implies that the effect of the disturbance d is eliminated. Similarly, the
map from reference r to output y is
GC
1 + GC
and if C 1 it can be seen that
GC
≈1
1 + GC
which implies that y = r, i.e., the output is tracking the reference closely. In
conclusion: Good tracking and disturbance rejection both require high loop gain.
2.1 Introduction
If the dynamic behavior of a physical system can be represented by an equa-
tion, or a set of equations, this is referred to as the mathematical model of
the system. Such models can be constructed from knowledge of the physical
characteristics of the system, i.e. mass for a mechanical system or resistance for
an electrical system. Because the systems under consideration are dynamic in
nature, the descriptive equations are usually differential equations. Differential
equations are often the initial description of a system. The variables are just
the inputs and outputs. If the differential equations can be linearized, then the
Laplace transform can be utilized to simplify the method of solution. a single-
input single-output process is described by its transfer function: the ratio of the
Laplace transform of output and input.
2.2.1 Definition
The Laplace transform of a function of time f (t), 0 ≤ t ≤ ∞ with f (t) = 0 for
t ≤ 0 is defined as Z ∞
F (s) = L [f (t)] = f (t)e−st dt (2.1)
0
11
12 CHAPTER 2. SYSTEM MODELLING
Example 2.1 Let f (t) be a unit step function defined as f (t) = 1 for t ≥ 0.
2.2.2 Properties
The application of the Laplace transform in many instances is simplified by
utilization of the properties of the transform. These properties are presented
here, for which no proofs are given.
Linearity
L[k1 f1 (t) ± k2 f2 (t)] = k1 F1 (s) ± k2 F2 (s)
Differentiation
df (t)
L = sF (s) − f (0)
dt
where f (0) is the limit of f (t) as t approaches 0. In general, for higher-order
derivatives of f (t),
n
d f (t)
L = sn F (s) − sn−1 f (0) − sn−2 f (1) (0) − · · · − f (n−1) (0)
dtn
where f (i) (0) denotes the ith-order derivative of f (t) with respect to t, evaluated
at t = 0.
Integration
Z t
F (s)
L f (τ )dτ =
0 s
Shift in time
Shift in frequency
L[e∓αt f (t)] = F (s ± α)
2.2. THE LAPLACE TRANSFORM 13
Covolution
L[f1 (t) ∗ f2 (t)] = F1 (s)F2 (s)
2.2.3 Theorems
Initial value theorem
A useful property of the Laplace transform known as the initial value theorem
which states that it is always possible to determine the initial value of the time
function f (t) from its Lapalce transform. We may state the theorem in this
way:
lim f (t) = lim sF (s)
t→0 s→∞
Find the Laplace transform of f (t) = cos ωt. [Linearity property] Example 2.3
d2 f
Find the Laplace transform of f (t) = . [Differentiation property] Example 2.4
dt2
Solution The Laplace transform is
df
F (s) = s2 F (s) − sf (0) −
dt t=0
14 CHAPTER 2. SYSTEM MODELLING
Example 2.5 Find the Laplace transform of f (t) = e−αt cos ωt. [Shifting in frequency]
where P (s) and Q(s) are polynomials of s. Rational functions are defined as the
ratio of two polynomials. It is assumed that the order of P (s) in s is greater
than that of Q(s), F (s) is said to be strictly proper. The polynomial P (s) may
be written
P (s) = sn + an−1 sn−1 + · · · + a1 s + a0
where a0 , a1 , · · · , an−1 are real coefficients. The roots of the polynomial P (s)
are referred to as poles of the function F (s).
The transfer function of a linear time-invariant system is the ratio of the Laplace
transform of the output to the Laplace transform of the input, with all initial
conditions assumed to be zero
Y (s)
H(s) =
X(s)
The transfer function of a system represents the relationship describing the dy-
namics of the system under consideration. Another way of defining the transfer
18 CHAPTER 2. SYSTEM MODELLING
function is to use the impulse response. The transfer function of a linear time-
invariant systems is defined as the Laplace transform of the impulse response,
with all initial conditions set to zero
Y (s)
H(s) = L[h(t)] =
X(s)
Example 2.9 Find the transfer function of the system described by the following differential
equation:
d2 y dy
+3 + 2y = 5
dt2 dt
with initial conditions
Solution Take Laplace transform and set all initial conditions to zero
5
s2 Y (s) + 3sY (s) + 2Y (s) =
s
5
Y (s) =
s(s2 + 3s + 2)
Y (s) 1
H(s) = = 2
(5/s) s + 3s + 2
v(t) = Ri(t)
2.4. MODELS OF ELECTRIC CIRCUITS 19
V (s) = RI(s)
For an inductor,
di(t)
v(t) = L
dt
Taking Laplace and assuming zero initial conditions
V (s) = sLI(s)
For a capacitor,
dv(t)
i(t) = C
dt
which transforms into the s-domain (assuming zero initial conditions) as
1
V (s) = I(s)
sC
The s-domain equivalents are shown in Figure 2.2.
Example 2.10 Determine the transfer function H(s) = Vo (s)/Io (s) for the circuit shown in
20 CHAPTER 2. SYSTEM MODELLING
Figure 2.3.
But
2(s + 4)Io
Vo = 2I2 = 1
s + 6 + 2s
Hence,
Vo (s) 4s(s + 4)
H(s) = = 2
Io (s) 2s + 12s + 1
Example 2.11 For the circuit shown in Figure 2.4, find the transfer function I2 (s)/V( s).
Solution The first step in the solution is to convert the network into Laplace
transform for impedance and circuit variables, assuming zero initial conditions,
as shown in Figure 2.5. The circuit requires two simultaneous equations to
solve for the transfer function. These equations can be found by summing volt-
ages around each mesh through which the assumed currents I1 (s) and I2 (s) flow.
Around Mesh 1,
R1 I1 (s) + LsI1 (s) − LsI2 (s) = V (s)
2.5. MODELS OF MECHANICAL SYSTEMS 21
Around Mesh 2,
1
LsI2 (s) + R2 I2 (s) + I2 (s) − LsI1 (s) = 0
Cs
By solving the two equations, we get
I2 (s) LCs2
H(s) = =
V (s) (R1 + R2 )LCs2 + (R1 R2 C + L)s + R1
as shown in Figure 2.6.
Mechanical systems require one or more differential equations, called the equa-
tions of motion, to describe it. We will begin by assuming a positive direction of
motion, for example, to the right. This assumed positive direction of motion is
similar to assuming a current direction in an electrical loop. Using our assumed
direction of positive motion, we first draw a free-body diagram, placing on the
body all forces that act on the body either in the direction of motion or opposite
to it. Next, we use Newton?s law to form a differential equation of motion by
summing the forces and setting the sum equal to zero. Finally, assuming zero
initial conditions, we take the Laplace transform of the differential equation,
separate the variables and arrive at the transfer function.
22 CHAPTER 2. SYSTEM MODELLING
Table 2.2: Force displacement translational relationship for spring, viscous damper
and mass.
Component Force-displacement
f (t) = Kx(t)
dx(t)
f (t) = B
dt
d2 x(t)
f (t) = M
dt2
Example 2.12 Find the transfer function X(s)/F( s) for the system shown in Figure 2.7(a).
1
M s2 + Bs + K
Figure 2.7: (a) Mass, spring and damper system; (b) block diagram.
2.5. MODELS OF MECHANICAL SYSTEMS 23
B B
Figure 2.8: (a) Free-body diagram of mass, spring and damper system; (b) trans-
formed free-body diagram (in s domain).
the differential equation of motion using Newton’s law to sum to zero all of the
forces shown on the mass in Figure 2.8(a):
d2 x(t) dx(t)
M 2
+B + Kx(t) = f (t)
dt dt
Taking the Laplace transform, assuming zero initial conditions,
which is represented in Figure 2.8(b). Solving for the transfer function yields
X(s) 1
H(s) = =
F (s) M s2 + Bs + K
The rotational system shown in Figure 2.9 consists of a disk mounted Example
on a 2.13
24 CHAPTER 2. SYSTEM MODELLING
Table 2.3: Torqur-angular displacement relationship for spring, viscous damper and
inertia.
shaft that is fixed at one end. The moment of inertia of the disk about the axis
of rotation is J. The edge of the disk is riding on the surface, and the viscous
friction coefficient between the two surfaces is B. The inertia of the shaft is
negligible, but the torsional spring constant is K. Assume the torque is applied
to the disk, as shown; then the torque or moment of equation about the axis of
the shaft is written from the free-body diagram of Figure 2.9(b) as
d2 θ(t) dθ(t)
T (t) = J 2
+B + Kθ(t) (2.2)
dt dt
e = Ke θ̇m (2.3)
where Ke is a constant of proprtionality called the back emf constant and θ̇m =
ωm is the angular velocity of the motor. Taking the Laplace transform we have
The relationship among the armature current ia , the applied armature voltage
va , and the back emf e is found by writing a loop equation around the Laplace
transformed armature circuit:
The torque developed by the motor is proprtional to the armature current, thus
To find the transfer function of the motor, we first substitute Equation (2.12)
and Equation (2.7) into Equation (2.13), yielding
θm (s) Km
=
Va (s) s [Ra (Jm s + b) + Km Ke ]
Km
=
s [Jm Ra s + bRa + Km Ke ]
Y (s) = G(s)E(s)
thus,
Y (s) = G(s)[R(s) − H(s)Y (s)]
Solving for Y (s), we obtain
Therefore, the transfer function relating the output Y (s) to the input R(s) is
Y (s) G(s)
=
R(s) 1 + G(s)H(s)
The gain of a single-loop negative feedback system is given by the forward gain
divided by the sum of one plus the loop gain. When the feedback is added in-
stead of subtracted, we call it positive feedback. In this case the gain is given by
the forward gain divided by the sum of 1 minus the loop gain.
A control system may have several feedback control loops as the one shown
in Figure 2.13. In principle, the block diagram of a closed-loop system, no mat-
ter how complicated it is, it can be reduced to the standard single loop form
shown in Figure 2.12. Reduction of complex block diagrams is facilitated by a
series of easily derivable transformations which are summarized in Table 2.4.
A block diagram of a multiple-loop feedback control system is shown in Figure Example 2.14
2.13. It is interesting to note that the feedback signal H1 (s)Y (s) is a positive
feedback signal, and the loop G3 (s)G4 (s)H1 (s) is a positive feedback loop. First,
to eliminate the minor loop G3 G4 H4 , we move H2 behind block G4 by using
rule 10 (see Table 2.4), and therefore obtain Figure 2.14.
Then eliminating the inner loop containing H2 /G4 , we obtain Figure 2.16(a).
Finally, by reducing the loop containing H3 , we obtain the closed-loop system
transfer function as shown in Figure 2.16(b).
30 CHAPTER 2. SYSTEM MODELLING
Example 2.15 Find the transfer function of the system shown in Figure 2.17.
Solution Moving the first summing point ahead of G1 , and the final take
off point beyond G4 gives a modified block diagram shown in Figure 2.18(a).
The block diagram in Figure 2.18(a) is then reduced to the form given in Figure
2.18(b).
G1 G2 G3 G4
Y (s) (1+G1 G2 H1 )(1+G3 G4 H2 )
= G 1 G 2 G 3 G 4 H3
R(s) 1+ (G1 G4 )(1+G1 G2 H1 )(1+G3 G4 H2 )
G1 G2 G3 G4
=
(1 + G1 G2 H1 )(1 + G3 G4 H2 ) + G2 G3 H3
2. Combining Y = P1 X ± P2 X
blocks in parallel
3. Removing a block Y = P1 X ± P2 X
from a forward
loop
4. Eliminating feed- Y = P1 (X ∓ P2 Y )
back loop
5. Removing a block Y = P1 (X ∓ P2 Y )
from a feedback
loop
6. Rearranging Z =W ±X ±Y
summing junc-
tions
7. Moving a sum- Z = PX ± Y
ming junction in
front of a block
32 CHAPTER 2. SYSTEM MODELLING
8. Moving a sum- Z = P (X ± Y )
ming junction be-
yond a block
9. Moving a takeoff Y = PX
point in front of a
block
3. Find the output response due to the chosen input action alone.
Find the complete output for the system shown in Figure 2.19 when both inputs Example 2.16
act simultaneously.
Y (s)
Solution The block diagram shown in Figure 2.19 can be reduced and sim-
plified to the form given in Figure 2.20.
Y (s)
Putting R2 (s) = 0 and replacing the summing point by +1 gives the block dia-
gram shown in Figure 2.21. In Figure 2.21 nothe that Y1 (s) is response to R1 (s)
acting alone.
Y1 (s)
R1 (s) 1+ G 1 G 2 H1
1+G2 H2
34 CHAPTER 2. SYSTEM MODELLING
or
G1 G2 R1
Y1 (s) =
1 + G2 H2 + G1 G2 H1
Now if R1 (s) = 0 and the summing point is replaced by −1, then the response
Y2 (s) to input R2 (s) acting alone is given by Figure 2.22. The choice as to
whether the summing point is replaced by +1 or −1 depends upon the sign at
the summing point.
Y2 (s)
Note that in Figure 2.22 there is a positive feedback loop. Hence the closed-loop
transfer functionn relating R2 (s) and Y2 (s) is
−G G H
1 2 1
Y2 (s)
= 1+G2 H2
R2 (s) 1 − −G 1 G2 H1
1+G2 H2
or
−G1 G2 H1 R2
Y2 (s) =
1 + G2 H2 + G1 G2 H1
Using the principle of superposition, the complete response is given by
or
G1 G2 R1 − G1 G2 H1 R2
Y (s) =
1 + G2 H2 + G1 G2 H1
points or nodes are used to represent variables. The nodes are connected by
line segments, called branches. A signal can transmit through a branch only in
the direction of the arrow. As an example consider a linear system represented
by a simple algebraic equation
y2 = a12 y1
where y1 is the input, y2 is the output, and a12 is the gain between the two
variables. The SFG representation is shown in Figure 2.23
y2 = a12 y1 + a32 y3
y3 = a23 y2 + a43 y4
y4 = a24 y2 + a34 y3 + a44 y4
y5 = a25 y2 + a45 y4
• The loop gain is the path gain of a feedback loop. For example, the loop
gain of the loop y2 − y4 − y3 − y2 in Figure 2.26 is a24 a43 a32 .
2.7. SYSTEM MODELLING DIAGRAMS 37
Mason’s Rule
Given an SFG or block diagram, the task of solving for the input-output rela-
tions by algebraic manipulation could be quite tedious. Fortunately, there is a
general gain formula available that allows the determination of the input-output
relations of an SFG by inspection.
Mason’s states that the input-output transfer function associated with a signal-
flow graph is given by P
Pk ∆k
G= k (2.11)
∆
where
∆ = 1 − L1 + L2 − L3 + · · · + (−1)m Lm
P P P P
and
Pk = gain of the k th forward path
L1 = gain of each closed loop in the graph
L2 = product of loop gains of any two nontouching loops (loops are called non-
touching if they have no node in common)
..
.
Lm = product of loop gains of any m nontouching loops
∆k = the value of ∆ remaining with the loops touching the path Pk are removed
4. Determine ∆ and ∆k .
5. Substitute all of the above information into the Mason’s gain formula.
Example 2.18 Determine the closed-loop transfer function Y (s)/R(s) of the SFG shown in
Figure 2.27.
Solution
1. Forward path: There is only one forward path between R(s) and Y (s),
and the forward-path gain is
P1 = G(s)
L1 : −G(s)H(s)
3. Non-touch loops: There are no non-touching loops since there is only one
loop.
4. ∆ and ∆1 : The forward path is in touch with the only loop. Thus, ∆1 = 1,
and
∆ = 1 − L1 = 1 + G(s)H(s)
P1 ∆ 1 G(s)
=
∆ 1 + G(s)H(s)
2.7. SYSTEM MODELLING DIAGRAMS 39
Determine the gain between y1 and y5 using the gain formula for the SFG shown Example 2.19
in Figure 2.25.
Solution
1. Forward path: There are three forward paths between y1 and y5 and
forward-path gains are
2. Closed loops: The four loops of the SFG are shown in Figure 2.26. The
loop gains are
hence X
L1 = a23 a32 + a34 a43 + a24 a43 a32 + a44
3. Non-touch loops: There is only one pair of non-touching loops; that is, the
two loops
y2 − y3 − y2 and y4 − y4
Thus the product of the gains of the two non-touching loops is
∆ = 1 − (a23 a32 + a34 a43 + a24 a43 a32 + a44 ) + a23 a32 a44
All the loops are in touch with forward paths P1 and P3 . Thus, ∆1 =
∆3 = 1. Two of the loops are not in touch with forward path P2 . These
loops are: y3 − y4 − y3 and y4 − y4 . Thus,
node, we simply connect a branch with unity gain from the existing node y2 to
a new node also designated as y2 , as shown in Figure 2.28.
Example 2.20 Determine the gain between y1 and y2 using the gain formula for the SFG
shown in Figure 2.28.
Solution
1. Forward path: There is only one forward path between y1 and y2 and
forward-path gain is
P1 = a12 Forward path: y1 − y2
2. Closed loops: The four loops of the SFG are shown in Figure 2.26. The
loop gains are
L1 : a23 a32 a34 a43 a24 a43 a32 a44
hence X
L1 = a23 a32 + a34 a43 + a24 a43 a32 + a44
3. Non-touch loops: There is only one pair of nontouching loops; that is, the
two loops
y2 − y3 − y2 and y4 − y4
Thus the product of the gains of the two nontouching loops is
L2 : a23 a32 a44
4. ∆ and ∆k : ∆ = 1 − (a23 a32 + a34 a43 + a24 a43 a32 + a44 ) + a23 a32 a44
Two of the loops are not in touch with forward path P1 . These loops
are: y3 − y4 − y3 and y4 − y4 . Thus,
∆1 = 1 − a34 a43 − a44
We can show that by including a source node (y1 in this case), we may write
y7 /y2 as
P
k Pk ∆k
∆
y7 y7 /y1
= = P from y1 to y7
y2 y2 /y1 k Pk ∆k
∆
from y1 to y2
Since ∆ is independent of the sources and sinks, the last equation is written
P
P ∆
k k k
y7 y7 /y1
= = from y1 to y7
y2 y2 /y1 P
k Pk ∆ k
from y1 to y2
Note that ∆ does not appear in the last equation. However, you must evaluate
it to be able to find ∆k .
Determine the gain between y7 and y2 for the SFG shown in Figure 2.29. Example 2.21
P1 = 1 Forward path: y1 − y2
y2 − y3 − y2 y4 − y5 − y4 y2 − y3 − y4 − y5 − y2 y4
with gains
L1 : −G1 H1 − G3 H2 − G1 G2 G3 H3 − H4
42 CHAPTER 2. SYSTEM MODELLING
Hence, X
L1 = −(G1 H1 + G3 H2 + G1 G2 G3 H3 + H4 )
3. Non-touch loops:
– Product of loop gains of any two nontouching loops (there are four
possible combinations), thus:
L2 : G1 G3 H1 H2 G1 H1 H4 G3 H2 H4 G1 G2 G3 H3 H4
and
X
L2 = G1 G3 H1 H2 + G1 H1 H4 + G3 H2 H4 + G1 G2 G3 H3 H4
2. Interconnect the nodes & indicate the directions of signal flow by using
arrows.
3. Identify the blocks → they are replaced with branches. For each negative
sum, a negative sign is included with the branch.
Example 2.22 Convert the block diagram in Figure 2.31 to a signal flow graph and determine
the transfer function using Mason’s gain formula.
Figure 2.31: (a) Block diagram a control system. (b) Equivalent signal-flow graph.
Solution
1. Forward path:
P1 = G1 G2 G3
P2 = G1 G4
2. Closed loops:
∆ = 1 + G1 G2 H1 + G2 G3 H2 + G1 G2 G3 + G1 G4 + G4 H2
5. Using (2.11), the transfer function between Y (s) and R(s) is written as
Y (s) P1 ∆ 1 + P2 ∆ 2 G1 G2 G3 + G1 G4
= =
R(s) ∆ ∆
2.7. SYSTEM MODELLING DIAGRAMS 45
Example 2.24 Construct a block diagram for a system described by the following set of equa-
tions
1
Pd (s) = Pc (s) (2.15)
CRd s + 1
1
Pi (s) = Pc (s) (2.16)
CRi s + 1
Pc (s) = KX(s) (2.17)
b a
X(s) = E(s) + Y (s) (2.18)
a+b a+b
A
Y (s) = [Pi (s) − Pd (s)] (2.19)
Ks
Solution Each dynamic equation represents a subsystem. Its block diagram
is constructed by a simple principle: treat the right hand side signals as the
input and the left hand side as the output.
3.1 Introduction
This chapter refers to the time-domain analysis of linear time-invariant con-
trol systems. The problem of time-domain analysis may be briefly stated as
follows: given the system (i.e., given a specific description of the system) and
its input, determine the time-domain behavior of the output of the system.
In the analysis problem, we will use selected input signals to test the response
of control systems. This response will be characterized by a selected set of
response measures. The basic motivation for system analysis is that one can
predict (theoretically) the system’s behavior.
(a) Transient response: Is that particular part of the response of the system
which tends to zero as time increases. It is a function only of the system
dynamics, and is independent of the input signal.
(b) Steady-state response: Is that particular part of the response of the sys-
tem that remains after the transient component has reached zero. It is a
function of both the system dynamics and the input signal.
49
50 CHAPTER 3. TIME-DOMAIN ANALYSIS OF CONTROL SYSTEMS
The total response of the system is always the sum of the transient and steady-
state components. In the design problem, specifications are usually given in
terms of the transient and steady-state performance, and controllers are de-
signed so that the specifications are all met by the design system.
1
δ (t) = , 0 ≤ t ≤ , >0
As → 0, δ (t) approaches the unit impulse δ(t). The major properties of this
function are
Z ∞
1. δ(t)dt = 1
0
Z ∞
2. g(t)δ(t − a)dt = g(a)
0
3. L[δ(t)] = 1
1
r(t) = 1, t≥0 =⇒ L[r(t)] =
s
1
r(t) = t, t≥0 =⇒ L[r(t)] =
s2
y(t)
K(1 − e−t/τ )
0.632K
Mathematically, the exponential term does not decay to zero in a finite length
of time. However, if the term continued to decay at its initial rate, it would
reach a value of zero in τ seconds. The parameter τ is called the time constant
and has the units of seconds. The exponential function decays to about 2% of
its initial value within 4 time constants. The output y(t) reaches about 63% of
its final vaue when t = τ .
Example 3.1 An example of a first order system is provided by the following RC circuit.
Vout (s) 1
= (K = 1, τ = RC)
Vin (s) 1 + sRC
Y (s) = G(s)R(s)
where G(s) is a given transfer function. From Section 2.2.3, the final-value
theorem of the Laplace transform is
exists, i.e., y(t) has a final value. For the case that the input is a unit step, R(s)
is equal to 1/s and
1
lim y(t) = lim sG(s) = lim G(s) (3.2)
t→∞ s→0 s s→0
= G(0)
G(0) is often called the dc gain of the system and is defined as the ratio of the
output of a system to a constant input after all transients has decayed. Care
must be taken to apply the Final value Theorem only to stable systems (i.e.,
y(t) is bounded) and with at most a single pole at s = 0.
Find the dc gain of the system whose transfer function is Example 3.2
3(s + 2)
G(s) =
(s2 + 2s + 10)
Solution Applying (3.2), we get
(3)(2)
dc gain = G(s) = = 0.6
s=0 (10)
ζ : damping ratio
K : dc gain
where p
s1 , s2 = −ζωn ± ωn ζ2 − 1
are the roots of the characteristic equation and are the poles of G(s).
ζ>1
ζ=1
are complex conjugate and have negative real parts. The real parts are zero if
ζ = 0 and the system is called undamped.
0≤ζ<1
Figure 3.7: Pole locations in the s-plane and the corresponding transient response
type.
56 CHAPTER 3. TIME-DOMAIN ANALYSIS OF CONTROL SYSTEMS
ωn2
Y (s) = G(s)R(s) =
s(s2 + 2ζωn s + ωn2 )
Expanding in partial fraction
K1 K2 s + K3
Y (s) = + 2
s s + 2ζωn s + ωn2
Using the cover-up method, we find
K1 = sY (s) =1
s=0
After equating the powers of s on the two sides of the above equation, we find
that K2 = −1 and K3 = −2ζωn , hence
1 s + 2ζωn
Y (s) = −
s s2 + 2ζωn s + ωn2
Completing the square
1 s + 2ζωn
Y (s) = −
s (s + ζωn )2 + ωn2 − ζ 2 ωn2
1 s + 2ζωn
= − 2
s
p
(s + ζωn )2 + ωn 1 − ζ 2
Using the trigonometric identity, sin(a + b) = sin a cos b + cos a sin b, (3.5) can
be written as
1
y(t) = 1 − p e−ζωn t sin(ωd t + α) (3.6)
1 − ζ2
3.5. RESPONSE OF SECOND ORDER SYSTEMS 57
where
1
p
−1 1 − ζ2
α = tan (3.7) p
ζ 1 − ζ2
p α
is as shown (note that cos α = ζ and sin α = 1 − ζ 2 ). The step responses for a
ζ
second-order system are shown in Figure 3.8 for several values of ζ as a function
of ωn t.
y(t)
ωn t
Figure 3.8: Step response for second-order system.
y(t) = 1 − cos ωn t
which clearly shows that no sinusoidal term exist. For the overdamped
case, we have two real poles at −ζωn ± ωd . The corresponding response
is easily obtained from
1 K2 K3
Y (s) = + +
s s + ζωn + ωd s + ζωn − ωd
as
y(t) = 1 + K2 e−(ζωn +ωd )t + K3 e−(ζωn −ωd )t
The effect of the characteristic equation roots on the damping of the second-
order system is further illustrated by Figure 3.9
Figure 3.9: Step response comparison for various pole locations in the s-plane.
3.6. SPECIFICATIONS OF A SECOND ORDER SYSTEM 59
The speed of the response is measured by the rise time, Tr , and the peak
time, Tp . For underdamped systems with an overshoot, the 0 − 100% rise time
is a useful index. If the system is overdamped, then the peak time is not defined,
and the 10 − 90% rise time, Tr1 , is normally used.
The tracking properties, i.e., the similarity with which the actual response
matches the step input is measured by the percent overshoot and settling
time, Ts . The % overshoot is defined as
Mp − yss
% overshoot = × 100% (3.8)
yss
60 CHAPTER 3. TIME-DOMAIN ANALYSIS OF CONTROL SYSTEMS
where Mp is the peak value of the time response, and yss is the steady state
value or the final value of the response.
The settling time, Ts , is defined as the time required for the system to set-
tle within a certain percentage, δ, of the input amplitude. This band of ±δ is
shown in Figure 3.10. In other words it is the time required for the transient
to decay to a small value so that y(t) is almost in the steady state. For second
order systems we usually determine the time, Ts , for which the response remains
within 2% of the final value. From Figure 3.11 we can determine an approximate
to Ts by computing the time when the decaying exponential e−ζωn t reaches 2%:
e−ζωn Ts = 0.02
or
ζωn Ts ≈ 4
Therefore, we have
4
Ts ≈ 4τ =
ζωn
The steady-state error of the system may be measured on the step response of
the system as shown in Figure 3.10. To obtain analytic expressions for Tp we
first differentiate (3.6) to obtain
ωn p
ẏ(t) = p e−ζωn t sin 1 − ζ 2 ωn t
1 − ζ2
and equating ẏ(t) = 0 gives
p
1 − ζ 2 ωn t = nπ, n = 1, 2, 3, · · ·
from which we get
nπ
t= p
ωn 1 − ζ 2
3.6. SPECIFICATIONS OF A SECOND ORDER SYSTEM 61
Thus the product ωn Tp is also a function of only ζ, and this is also plotted in
Figure 3.12. Since Tp is an approximate indication of the rise time, Figure 3.12
also roughly indicates rise time. As ζ increases from 0 to 1, Mp , Ts , and the
percent overshoot decreases, while Tp and Tr increase.
Figure 3.12: Percent overshoot and normalized peak time versus damping ratio ζ.
62 CHAPTER 3. TIME-DOMAIN ANALYSIS OF CONTROL SYSTEMS
Comment
The transient response of the system may be described in terms of two factors:
1. The speed of response, measured by the rise time, Tr , and the peak
time, Tp .
2. The tracking properties, i.e., the closeness of the response to the desired
response, is measured by the % overshoot and the settling time, Ts .
It is important to realize that the two factors are contradictory requirements;
thus, a compromise must be obtained.
Solution We have
• ωn is the radial distance from the roots to the origin of the s-plane.
• ζωn is the real part of the roots.
• ωd is the imaginary part of the roots.
• ζ is the cosine of the angle between the radial line to the roots and the
negative axis when the roots are in the left-halp s-plane, or ζ = cos α.
The effect of increasing ωn and ζ on pole locations in the s-plane is shown in
Figure 3.14.
Figure 3.14: Pole locations tin the s-plane as ωn and ζ increases respectively.
Faster response
design the settling is specified to be less than or equal to some value Ts,desired ,
64 CHAPTER 3. TIME-DOMAIN ANALYSIS OF CONTROL SYSTEMS
ζωn ≥ 4/Ts,desired , and the pole locations are then restricted to the region of
the s-plane indicated in Figure 3.15. Hence the speed of response is increased
by moving the poles to the left in the s-plane.
Decreasing the angle α reduces the percent overshoot. Hence, specifying the
percent overshoot to be less than a particular value restricts the pole locations
to the region of the s-planre, as shown in Figure 3.16
Lines of constant Tp , % overshoot, and Ts are shown in Figure 3.17. Note that
Ts2 < Ts1 ; Tp2 < Tp1 ; and %OS1 < %OS2 .
Suppose that, in the design of a second-order system, the %overshoot in a step Example 3.4
response is limited to 4.32%. A maximum settling time of 2s is also required.
On an s-plane show the region to which the pole locations are limited to.
Consider the pole plot shown in Figure 3.19. Find ζ, ωn , Tp , and Ts . Example 3.5
66 CHAPTER 3. TIME-DOMAIN ANALYSIS OF CONTROL SYSTEMS
√
The natural frequency ωn is given by ωn = 72 + 32 = 7.616 rad/s
π π
The peak time Tp is Tp = = = 0.449s
ωd 7
4 4
The settling time Ts is Ts = = = 1.33s
ζωn 3
Figure 3.20: Step response as poles move with constant real part.
In Figure 3.21 the step responses are shown as the poles are moved in horizontal
3.8. STEADY-STATE ACCURACY 67
direction, keeping the imaginary part the same. As the poles move to the left,
the response damps out rapidly, while the frequency remains the same. Notice
that the peak time is the same for all waveforms.
Figure 3.21: Step response as poles move with constant imaginary part.
In Figure 3.22 the poles are moved along a constant radial line. We see that the
% overshoot remains the same. The farther the poles are from origin, the more
rapid the response.
Figure 3.22: Step response as poles move with constant damping ratio.
The output Y (s) is required to track the reference input R(s). The input into
plant G(s) is the tracking error
To calculate the steady-state error, we utilize the final value theorem, as long
as E(s) does not have any poles in the right half of the s-plane, except maybe,
at s = 0, then
sR(s)
ess = lim e(t) = lim sE(s) = lim
t→∞ s→0 s→0 1 + G(s)
Step Input The steady-state error for a step input is
sR(s) 1
ess = lim =
s→0 1 + G(s) 1 + G(0)
Note that in this case, the steady-state error is determined by the dc gain of
G(s). The larger is dc gain, the smaller is the steady-state error. Furthermore,
if G(s) has one or more poles at s = 0, then lims→0 G(s) = ∞. In this case,
ess = 0. A pole at s = 0 implies G(s) includes an integrator. The number of
poles at the origin of the loop gain (i.e. the number of integrators)
defines the system’s type number, N.
1
1
Ramp Input Now consider the steady -state error in response to a ramp input
r(t) = t. The Laplace transform of the ramp input is R(s) = 1/s2 and
s(1/s2 ) 1 1
ess = lim = lim = lim
s→0 1 + G(s) s→0 s + sG(s) s→0 sG(s)
Again, the steady-state error depends upon the number of integrators, N . For
a type zero system, N = 0, the steady-state error is infinite. For a type one
system, N = 1, the error is
1
ess =
Kv
where Kv the velocity error constant and is computed as
Kv = lim sG(s)
s→0
For N ≥ 2, Kv is infinite and the steady-state error for a ramp input is zero.
Parabolic Input for a parabolic input r(t) = t2 /2, we take R(s) = 1/s3 ,
the steady-state error is
s(1/s3 ) 1
ess = lim = lim 2
s→0 1 + G(s) s→0 s G(s)
The steady-sate error is infinite for type zero and type one systems. For type
two, N = 2, and we obtain
1
ess =
Ka
70 CHAPTER 3. TIME-DOMAIN ANALYSIS OF CONTROL SYSTEMS
The error constants and the steady-state error for the three inputs are summa-
rized in Table 3.1.
1 1 1
Type R(s) = R(s) = R(s) = Error constants
s s2 s3
1
0 ess = ∞ ∞ Kp = lim G(s)
1 + Kp s→0
1
1 ess = 0 ∞ Kv = lim sG(s)
Kv s→0
1
2 ess = 0 0 Ka = lim s2 G(s)
Ka s→0
200(s + 1)2
G(s) =
(s + 2)(s + 3)(s + 4)
Find the steady-state error when (a) r(t) is a unit step, (b) r(t) is a unit ramp.
Solution G(s) is a type zero system, therefore for a step input, the posi-
tion error constant, Kp = G(0) = 200/24 = 8.333. Then
1 1
ess = = = 0.1071
1 + Kp 9.333
3.8. STEADY-STATE ACCURACY 71
1
ess = =∞
Kv
72 CHAPTER 3. TIME-DOMAIN ANALYSIS OF CONTROL SYSTEMS
Chapter 4
Stability Analysis
Stability is the most crucial issue in designing any control system. One of the
most common control problems is the design of a closed loop system such that
its output follows its input as closely as possible. If the system is unstable such
behavior is not guaranteed. Unstable systems exhibit an unbounded output,
i.e., a response blowing up to infinity as time increases. This usually cause the
system to suffer serious damage such as burn out, break down or it may even
explode. Therefore, for such reasons our primary goal is to guarantee stability.
As soon as stability is achieved one seeks to satisfy other design requirements,
such as speed of response, settling time, steady state error, etc.
4.1 Introduction
To help make the later mathematical treatment of stability more intuitive let
us begin with a general discussion of stability concepts and equilibrium points.
Consider the ball which is free to roll on the surface shown in Figure 4.1. The
ball could be made to rest at points A, E, F , and G and anywhere between
points B and D, such as at C. Each of these points is an equilibrium point of
the system.
A small perturbation away from points A or F will cause the ball to diverge
from these points. This behavior justifies labeling points A and F as unstable
73
74 CHAPTER 4. STABILITY ANALYSIS
equilibrium points. After small perturbations away from E and G, the ball will
eventually return to rest at these points. Thus E and G are labeled as stable
equilibrium points. If the ball is displaced slightly from point C, it will normally
stay at the new position. Points like C are sometimes said to be neutrally stable.
Stability deals with the following questions. If at time t0 the system is per-
turbed from its equilibrium point, does the system return to that point, or
remain close to it, or diverge from it?
Example 4.1 For an LTI system with impulse response h(t) = e−3t u(t), determine the stabil-
ity of this causal LTI system.
bm sm + bm−1 sm−1 + · · · + b1 s + b0
G(s) =
an sn + an−1 sn−1 + · · · + a1 s + a0
where the denominator polynomial is called the characteristic polynomial of
G(s). The equation
Assume that the roots pi of the characteristic equation are real or complex,
but are distinct. The solution to the differential equation whose characteristic
equation is given by (4.2) may be written using partial-fraction expansion as
n
X
y(t) = Ki epi t (4.3)
i=1
where pi are the roots of (4.2). The system is stable if and only if every term
in (4.3) goes to zero as t → ∞:
e pi t → 0 for all pi
This will happen if all the poles of the system are strictly in the LHP, where
Re{pi } < 0
If any poles are repeated, the response must be changed from that of (4.3)
by including a polunomial in t in place of Ki , but the conclusion is the same.
Figure 4.2: Time functions associated with pole locations in the s-plane.
76 CHAPTER 4. STABILITY ANALYSIS
The following examples illustrates the stability conditions of systems with ref-
erence to the poles of the transfer function G(s).
20
G(s) = Stable
(s + 1)(s + 2)(s + 3)
20(s + 1)
G(s) = Unstable due to the pole at s = 1
(s − 1)(s2 + 2s + 2)
20(s − 1)
G(s) = Marginally stable due to s = ±j2
(s + 2)(s2 + 4)
10
G(s) = Unstable due to the multiple poles
(s2 + 4)2 (s + 10) at s = ±j2
range of values of a parameter that ensures stability. For example in the spring-
mass-damper system the characteristic equation is M s2 + Bs + K, one can not
determine the ranges of M, B and K to ensure stability.
Consider the second order polynomial, assuming all coefficients are real
a1 = −(p1 + p2 ) and a0 = p1 p2
If p1 and p2 are stable, we have a1 > 0 and a0 > 0. Consider next the third
order polynomial
a2 = −(p1 + p2 + p3 )
a1 = (p1 p2 + p2 p3 + p1 p3 )
a0 = −p1 p2 p3
Again if all roots are stable, all the coefficients will have the same sign. How-
ever, this condition is not sufficient, for it is quite possible that an
equation with all its coefficients nonzero and of the same sign still
will not have all the roots in the left half of the s-plane. Consider for
example the polynomial s3 + s2 + 2s + 8, clearly all the coefficients have the
same sign however not all roots are in the LHP.
We conclude
• If all roots are stable, all the polynomial coefficients will be positive.
and work with P̂ (s) instead. Note that in the case a0 = 0, P (s) will have at
least one root at the origin and we conclude that the LTI system is marginally
stable or unstable.
where
1 an an−2 1 an an−4
b1 = −
an−1 b2 = −
an−1 an−3 an−1 an−1 an−5
1 an−1 an−3 1 an−1 an−5
c1 = − c2 = −
b1 b1 b2 b1 b1 b3
..
.
1 k1 k2
m1 = −
l1 l1 0
Once the Routh’s array has been completed, we investigate the signs of the
coefficients in the first column of the array. The roots of the equation are all in
the left half of the s-plane if all the elements of the first column of the Routh’s
array are of the same sign. The number of unstable roots is equal to the number
of sign changes in the first column of the array.
P (s) = s3 − s2 + s + 6
This equation has one negative coefficient. Thus, we know without applying
Routh’s test that not all the roots of the equation are in the LHP.
P (s) = s3 + s2 + 2s + 8
s3 1 2
2
s 1 8
s -6
1 8
The two sign changes (from +1 to -6 and from -6 to +8) indicates two unstable
roots.
80 CHAPTER 4. STABILITY ANALYSIS
1. The first element of a row is zero, with at least one nonzero element in the
same row, the procedure is modified by replacing that first element with a
small number such that || 1 and proceeding as before.
2. Every entry in a row is zero, the last modification will not give useful
information and another modification is needed.
Since the first element of the s3 row is zero, the elements in the s2 row would
all be infinite. To overcome this difficulty, we replace the zero in the s3 row by
a small positive number and then proceed with the array
s5 1 2 11
4
s 2 4 10
3
s 0→ 6 0
There are two sign changes (irrespective of the sign of ) indicating two unstable
roots.
Special Case: Zero rows. If all the coefficients in a row are zero, a pair
of roots of equal magnitude and opposite sign is indicated. These could be two
real roots with equal magnitudes and opposite signs or two conjugate imaginary
4.3. ROUTH-HURWITZ STABILITY CRITERION 81
roots. The zero row is replaced by taking the coefficients of dPa (s)/ds, where
Pa (s), called the auxiliary polynomial, is obtained from the values in the row
above the zero row. Why? A zero row implies that a polynomial Pa (s) has only
even or odd powers. It turns out in this case, Pa (s) and Pa (s) + dPa (s)/ds have
exactly the same number of RHP poles (proof beyond the scope of theis course).
As the goal is just to find the number of RHP poles, we can use dPa (s)/ds as a
surrogate to continue the procedure. The pair of roots can be found by solving
dPa (s)/ds = 0. The roots of Pa (s) are also the roots of the the ploynomial P (s).
The following example combines case 1 and case 2 problems: polynomial: Example 4.6
P (s) = s4 + 4
The Routh array is then
s4 1 0 4 ←− Pa (s) = s4 + 4
dPa (s)
s3 0→4 0 0 ←− Coefficients of = 4s3
ds
s2 0→ 4
s −16/
1 4
The two sign changes indicate two unstable roots.
In Summary, the three cases that occur in the application of the Routh-
Hurwitx criterion are as follows:
82 CHAPTER 4. STABILITY ANALYSIS
Case 1. No elements in the first column are zero. There are no problems
in completing the array.
Case 2. There is at least one nonzero element in a row, with the first
element equal to zero. This always indicates an unstable system. The
first element (which is zero) is replaced with the value , 1, and the
calculation of the array continues.
Case 3. All elements in a row are zero. This always indicates a system
that is not stable, but it may be marginally stable. This case can be
analyzed through the use of the auxiliary equation, as described earlier. If
the system is marginally stable the roots on the jω axis are also the roots
of the auxiliary equation.
K(s) G(s)
1 + G(s)K(s) = 0
Suppose we write
Ng (s) Nk (s)
G(s) = and K(s) =
Dg (s) Dk (s)
where Ng (s), Dg (s), Nk (s), and Dk (s) are all polynomials. Then, the closed
loop transfer function is given
Ng (s)Nk (s)
Dg (s)Dk (s)
H(s) =
Ng (s)Nk (s)
1+
Dg (s)Dk (s)
Ng (s)Nk (s)
=
Ng (s)Nk (s) + Dg (s)Dk (s)
4.4. APPLICATIONS IN FEEDBACK DESIGN 83
The poles of the closed-loop system are also given by the roots of the charac-
teristic polynomial
Ng (s)Nk (s) + Dg (s)Dk (s)
The poles of the open-loop system G(s) are the roots of its characteristic poly-
nomial
Dg (s) = 0
which are generally different from the closed-loop poles.
1
G(s) =
s3 + 5s2 + 2s − 8
Let K(s) = K be a constant controller. Find the range of values of K for which
the closed-loop is stable.
1
G(s) =
s3 − s2 − 10s − 8
Find the range of values of K for which the closed-loop is stable.
which implies
s3 − s2 − 10s + (K − 8) = 0
It follows that no choice of K can ensure all coefficients have the same sign. We
conclude that G(s) cannot be stabilized by a constant controller and dynamic
compensation is need.
Chapter 5
Root-Locus Analysis and Design
In this chapter we introduce one of the major analysis and design methods
discussed in this course. The method is the root-locus procedure; it indicates
to us the characteristics of a control system’s transient response. We have seen
that the response of an LTI system is largely determined by the location of its
poles.
1
K
s(s + 2)
s2 + 2s + k = 0
We see, since the polynomial is second order, that the system is stable for all
positive values of K. It is not evident for this example exactly how the value of
85
86 CHAPTER 5. ROOT-LOCUS ANALYSIS AND DESIGN
1 s2 + 2s + 1 = 0 s = −1 ± j0
2 s2 + 2s + 2 = 0 s = −1 ± j
K affects the transient response. Table 5.2 show the roots of the characteristic
equation for different values of K. To investigate some of the effects of choosing
different values of K, we plot the roots of the system characteristic equation in
the s-plane. These roots are plotted in Figure 5.2 for 0 ≤ K ≤ ∞.
We can see from the plot that for 0 < K < 1, the roots are real with different
time constants. For K = 1, the roots are real and equal, and the system is
critically damped. For K > 1, the roots are complex with a time constant of
1s, with the value of ζ decreasing as K increases. Hence, as K increases with
the roots being complex, the overshoot in the transient response increases.
The plot in Figure 5.2 is called the root locus of the system in Figure 5.1.
The root-locus of a system is a plot of the roots of the system char-
acteristic equation (closed-loop poles) as K varies from 0 to ∞.
For an nth order system, the root-locus is a family of n curves traced out by the
n closed-loop poles as K is varied from zero to infinity. Plotting the root locus
for negative values of K will be considered later.
5.1. ROOT-LOCUS PRINCIPLES 87
K G(s)
KG(s) = 180◦
and in general
G(s) = ±r(180◦ ) r = 1, 3, 5, · · ·
Equation (5.6) called the angle criterion may be interpreted as follows: For
a point s1 to be on the root-locus, the sum of all angles for vectors between
open-loop poles and zeros to point s1 must be equal 180◦ . The angle criterion is
illustrated in Figure 5.4 for the function
s − z1
G(s) =
(s − p1 )(s − p2 )
In Figure 5.4 the poles of G(s) are marked × and the zero is marked
. Suppose
that the point s1 is to be tested to determine if it is is on the root-locus. For
this point to be on the locus, we must have G(s1 ) = ±180◦ or equivalently
The angle from the zero term s−z1 can be computed by drawing a line from the
location of the zero at z1 to the test point s1 . In this case the line has a phase
angle marked θ1 on Figure 5.4. In a similar fashion, the vector from the pole
s = p1 to the test point s1 is shown with angle θ2 , and the angle of the vector
from the pole s = p2 to s1 is shown with angle θ3 . Thus the angle condition
(5.6) becomes
θ1 − θ2 − θ3 = ±r(180◦ )
for the point s1 to be on the root-locus.
In general the condition for a point in the s-plane to be on the root-locus is that
X X
(angles form zi ) − (angles from pi ) = ±r(180◦ ) r = 1, 3, 5, · · ·
i i
Check whether the point s0 = −1 + 2j lies on the root-locus for some value of Example 5.2
K if
s+1
G(s) =
s[(s + 2)2 + 4](s + 5)
Solution For s0 to be on the locus, we must have G(s0 ) = ±180◦ . Therefore,
G(s0 ) = (s0 + 1) − s0 − [(s0 + 2)2 + 4] − (s0 + 5)
RULE 2. The root-locus includes all points on the real axis to the
left of an odd number of poles and zeros.
This follows from the angle criterion, we consider first that all poles and zeros
of the open-loop transfer function are on the real axis, and we test points on
the real axis to determine if these points are on the locus.
Consider an open-loop transfer function G(s) of two poles and one zero as
illustrated in Figure 5.7. If we take a test point s on the real axis to the right of
the zero z1 as shown in Figure 5.7(a) we find that G(s) = 0. Hence, the angle
criterion is not satisfied, and we can see that any point to the right of the zero
z1 cannot be on the root locus.
Consider now a point s between the zero z1 and the pole p1 . In this case
5.2. RULES & STEPS FOR PLOTTING THE ROOT-LOCUS 91
− z1 − s − p 1 − s − p 2
(a) G(s)= |s {z } | {z } | {z }
0 0 0
− z1 − s − p1 − s − p2
(b) G(s)= |s {z } | {z } | {z }
180 0 0
− z1 − s − p1 − s − p2
(c) G(s)= |s {z } | {z } | {z }
180 180 0
− z1 − s − p1 − s − p2
(d) G(s)= |s {z } | {z } | {z }
180 180 180
G(s) = 180◦ , as shown in Figure 5.7(b). However the angles from the poles
p1 and p2 are still 0◦ . Thus the angle requirement is satisfied, and any point
between z1 and p1 is on the locus.
For a point s between p1 and p2 , (see Figure 5.7(c)) the angle from z1 is still
180◦ , as now is the angle from p1 . The angle from p2 is still 0◦ ; hence the angle
requirement is not satisfied and no points between p1 and p2 are on the locus.
If the point s is to the left of the pole p2 , the angles from z1 , p1 , and p2 are all
180◦ , and the angle criterion is satisfied as shown in Figure 5.7(d).
For the case that we have complex poles or zeros, the preceding discussion
still applies. For example, two complex conjugate poles are shown in Figure 5.8.
Since complex poles (and zeros) must occur in conjugate pairs, the sum of the
angles from a pair of poles (or zeros) to a point on the real axis will always be
0◦ (or 360◦ ). Hence complex poles and zeros do not affect the part of the root
locus that lies on the real axis.
Summary:
Figure 5.9: The real axis parts of the locus are to the left of an odd number of poles
and zeros.
n(s) (s − z1 )(s − z2 ) · · · (s − zm )
G(s) = = where n>m
d(s) (s − p1 )(s − p2 ) · · · (s − pn )
sm + b1 sm−1 + · · · + bm
=
sn + aa sn−1 + · · · + an
We assume that n(s) and d(s) are monic polynomials (monic means the coef-
ficient of the highest power of s is 1). The closed-loop characteristic equation
5.2. RULES & STEPS FOR PLOTTING THE ROOT-LOCUS 93
can be written as
n(s) 1 n(s)
1+K = 0 ⇐⇒ d(s) + Kn(s) = 0 ⇐⇒ + =0 (5.7)
d(s) K d(s)
If K = 0, then from (5.7) implies d(s) = 0 (i.e., the poles of G(s)). Therefore,
for K = 0 the roots of the closed-loop characteristic equation 1 + KG(s) = 0 are
the open-loop poles. The points of the root-locus where K = 0 are sometimes
called the starting or departure points of the root-locus.
Note from the above discussion that we have n poles and m zeros. If m of
the n poles will terminate at m zeros, where will the n − m poles terminate. As
Rule 3 states they will terminate at infinity, the question remains which infinity,
Rule 4 next clarifies the matter.
Recall from the discussion of Rule 3 for K → ∞, G(s) = 0 if n(s) is zero for
a finite s. The root locus will approach the open-loop zeros. To see a second
manner in which G(s) may go to zero, we express the characteristic equation
1 + KG(s) = 0 as
sm + b1 sm−1 + · · · + bm
1+K =0 (5.8)
sn + aa sn−1 + · · · + an
Since n > m, it is clear that G(s) goes to zero as s → ∞. In fact, for very large
values of s (5.8) can be approximated by
1
1+K =0 (5.9)
(s − σA )n−m
To see why (5.9) is a good approximation to (5.8), try to imagine what would
we see if we could observe the locations of poles and zeros from a distance point
near infinity: They would appear to cluster near the s-plane origin as shown in
Figure 5.10(a). Thus m zeros would cancel the effects of m poles, and the other
n − m poles would appear to be in the same place, namely at s = σA as shown
in Figure 5.10(b). If α = n − m we may write (5.9) as
1
1+K =0
(s − σA )α
94 CHAPTER 5. ROOT-LOCUS ANALYSIS AND DESIGN
−σA
n − m Poles
(a) (b)
We say that the locus of (5.8) is asymptotic to the locus of (5.9) for large values
of K and s.
To find the locus, no matter how far away the point s is on the s-plane it
must satisfy the angle criterion. Since all α poles appear to be in the same
place the angle condition gives
αφA = ±r(180◦ ) r = 1, 3, 5, · · ·
(180◦ )
=⇒ φA = ±r
α
The angles φA are the angles of asymptotes of the root-locus. Table 5.2 gives
these angles for small values of α.
α Angles
0 No asymptotes
1 180◦
2 ±90◦
3 ±60◦ , 180◦
4 ±45◦ , ±135◦
Finding σA
To determine σA we make use of polynomial properties discussed in Section
4.3.1. Write G(s) as
(s − z1 )(s − z2 ) · · · (s − zm )
G(s) =
(s − p1 )(s − p2 ) · · · (s − pn )
m
!
X
m
s − zi sm−1 + · · ·
i=1
= n
!
X
sn − pi sn−1 + · · ·
i=1
Dividing both the numerator and the denominator by the numerator gives
1
G(s) = n m
! (5.10)
X X
sn−m − pi − zi sn−m−1 + · · ·
i=1 i=1
Hence, X X
pi −
zi
σA =
n−m
P P
Notice that in the sum pi and zi the imaginary parts always add to zero
since complex poles and zeros always occur in complex conjugate pairs.
In summary the loci proceed to the zeros at infinity along asymptotes centered
at σA and with angles φA . when the number of m finite zeros is less than the n
number of poles, then n − m sections of loci must end at zeros at infinity. these
sections of loci proceed to the zeros at infinity along asymptotes as k approaches
infinity. These linear asymptotes are centered at the point σA on the real axis.
96 CHAPTER 5. ROOT-LOCUS ANALYSIS AND DESIGN
The most important use of this rule is to compute the angle of departure from a
complex pole. This angle of departure can sometimes be an aid in determining
the final shape of the root locus. To illustrate this rule consider the poles and
the zero shown in Figure 5.12. The vector angles at one complex pole p1 is also
shown in Figure 5.12. The radius of the circle around the pole p1 is actually very
small in relation to the distance to the zero and the other pole. The angles at a
test point s0 , an infinitesimal distance from p1 , must meet the angle criterion.
Therefore,
α − θ1 − θ2 = 180◦
5.2. RULES & STEPS FOR PLOTTING THE ROOT-LOCUS 97
s0
p1
First we take a test point s0 very near pole 2 at −4 + 4j and compute the
angle of G(s0 ). This situation is sketched in Figure 5.13. We select the test
point close enough to pole 2 that the angles φ1 and φ3 to the test point can be
98 CHAPTER 5. ROOT-LOCUS ANALYSIS AND DESIGN
considered the same as those angles to pole 2. Thus φ1 = 90◦ and φ3 = 135◦ ,
and φ2 can be calculated from the angle condition
as shown in Figure 5.14. By the complex conjugate symmetry of the plots, the
angle of departure of the locus near pole 1 at −4 − 4j will be +45◦ . Note for
a multiple pole of order q we must count the angle from the pole q times.
STEP 5. Estimate (or compute) the points where the locus crosses
the imaginary axis.
The points where the root-locus intersect the imaginary axis of the s-plane, and
the corresponding values of K, may be determined by means of the Routh-
Hurwitz criterion. For the third-order example we are using, the characteristic
equation is
K
1+ =0
s[(s + 4)2 + 16]
which is equivalent to
s3 + 8s2 + 32s + K = 0
s3 1 32
2
s 8 K
256 − K
s1 0
8
1 K
In this case we see that the s1 row coefficients are all zeros when K = 256
indicating a root on the imaginary axis. Thus K = 256 must correspond to
a solution at s = jω0 for some ω0 . Substituting this data into the auxiliary
equation gives us
−8ω02 + 256 = 0
√
Clearly the solution is ω0 = ± 32 = ±5.66, which is plotted in Figure 5.15.
Breakaway at s = −1
−2 −1 0
Maximum gain
at s = −1
−2 0
complex roots. Notice in Figure 5.16 that the gain K, as a function of the real
roots s, must have a local maximum at the breakaway points, so that, with
1
K=−
G(s)
and s considered a real variable, we require
dK
=0 (5.13)
ds
If we express G(s) as a ratio of two polynomials n(s) and d(s) the above equation
can be written as
dK d 1 d d(s)
= − = − =0
ds ds G(s) ds n(s)
The differentiation with respect to s, yields
d d(s) 1 d 1 d
− = − d(s)(−1) 2 [n(s)] + [d(s)]
ds n(s) n (s) ds n(s) ds
Equating the right hand side of the equation above to zero implies
d d
n(s) [d(s)] − d(s) [n(s)] = 0
ds ds
It is important to point out that the condition for a breakaway point given in
(5.13) is necessary but not sufficient. In other words, all breakaway points on
5.2. RULES & STEPS FOR PLOTTING THE ROOT-LOCUS 101
the root-locus must satisfy (5.13), but not all solutions of (5.13) are breakaway
points.
d d(s)
d(s) = s3 + 8s2 + 32s = 3s2 + 16s + 32
ds
the points of possible multiple roots or breakaway are given by
3s2 + 16s + 32 = 0
or
s0 = −2.67 ± 1.89j
The breakaway point must be real and lies on the root-locus, hence, for this
example there is no breakaway point.
5.3 Examples
In all the examples to follow, G(s) is the transfer function of a system to be
controlled using constant-gain in the forward path, with K ≥ 0.
STEPS 1 and 2. Mark the poles and zeros on the s-palne and draw the real axis
portion of the locus:
σA
STEP 4. We compute the departure angles from the poles. We draw a small
circle around the two poles at s = 0. The angles from the zero at −1 and from
the pole at −10 are both zero, and the angles from the two poles at the origin
are the same. Therefore, the root locus condition is
STEP 5. We compute the points where the locus crosses the imaginary axis:
s+1
1+K =0
s2 (s + 9)
s3 + 9s2 + Ks + K = 0
STEP 6. We locate the points of multiple roots, which will include breakaway
and break-in points:
d n(s)
n(s) = s + 1 =1
ds
d d(s)
d(s) = s3 + 9s2 = 3s2 + 18s
ds
The possible multiple roots are at
The points of multiple roots are on the locus, but we have repeated roots in the
derivative, which indicates that we have three roots at the same place. Note we
can apply the rule of departure angles to the triple root at s = −3, we find that
(s + 1)
Figure 5.18: Root locus for G(s) = .
s2 (s + 9)
STEPS 1 and 2. Mark the poles and zeros on the s-palne and draw the real axis
portion of the locus:
We can observe at once that, along the line s = −1 + jω the angle criterion is
always satisfied. This is a special case since the angles of the real axis poles to
any point on this line will form an isosceles triangle and always add to 180◦ .
s4 1 9 K
3
s 4 10
s2 6.5 K
65 − 4K
s1 0
6.5
1 K
In this case we see that the s1 row coefficients are all zeros when K = 16.25
indicating a root on the imaginary axis. Thus K = 16.25 must correspond to
a solution at s = jω0 for some ω0 . Substituting this data into the auxiliary
equation gives us
−6.5ω02 + 16.25 = 0
√
Clearly the solution is ω0 = ± 2.5 = ±1.58.
106 CHAPTER 5. ROOT-LOCUS ANALYSIS AND DESIGN
d n(s)
n(s) = 1 =0
ds
d d(s)
d(s) = s4 + 4s3 + 9s2 + 10s = 4s3 + 12s2 + 18s + 10
ds
The quadratic has roots −1 ± 1.22j. Since these points are on the line between
the complex poles, they are points of multiple roots on the locus.
STEP 7. The complete sketch is given in Figure 5.19. Notice that we have
complex multiple roots. Branches of the locus come together at −1 ± 1.22j and
break away at 0◦ and 180◦ .
Sketching root loci relies heavily on experience. Figure 5.20 gives several loci
for low-order systems; these should be studied to familiarize yourself with some
of the characteristics of root loci.
1
Figure 5.19: Root locus for G(s) = .
s(s + 2)[(s + 1)2 + 4]
5.3. EXAMPLES 107
dG(s)
=0
ds
Solution The root locus and the complementary root locus are shown in Fig-
ure 5.21. In the complementary root-locus, the locus on the real axis occurs to
the left of an even count of poles and zeros. Since zero is considered even, root
locus on the real axis will occur only to the right of the zero at the origin. The
break-in points are s = ±2.06. Thus the break-in point for the complementary
root locus is at s = 2.06. After the break-in, one closed-loop pole migrates to
the zero at the origin and the other to the right toward infinity.
5.4. THE COMPLEMENTARY ROOT LOCUS 109
s
Figure 5.21: Root locus for G(s) = , −∞ < K < ∞.
(s − 0.5 − 2j)(s − 0.5 + 2j)
Solution The complementary root locus on the real axis occurs to the right
of the pole at the origin, between the zero at s = −1 and the pole at s = −4,
and to the left of the pole at s = −10. The number of zeros at infinity, α = 2, so
there are two asymptotes at 0◦ and 180◦ . The root locus and the complementary
root locus are shown in Figure 5.22.
s+1
Figure 5.22: Root locus for G(s) = , −∞ < K < ∞.
s(s + 4)(s + 10)
110 CHAPTER 5. ROOT-LOCUS ANALYSIS AND DESIGN
1
G(s) =
s[(s + 4)2 + 16]
For this transfer function, the locus was plotted in Figure 5.17 and is repeated
here in Figure 5.23. On Figure 5.23 the lines corresponding to a damping ratio
of ζ = 0.5 are sketched, and the points where the locus crosses these lines are
marked with (•). Suppose we wish to set the gain so that the poles are located
at the dots. This corresponds to selecting the gain so that two of the closed-loop
poles have a damping ratio of ζ = 0.5. What is the value of K when a root is
1
Figure 5.23: Root locus for G(s) = showing calculations of gain K.
s[(s + 4)2 + 16]
5.6. DYNAMIC COMPENSATION 111
available, only two such techniques that have been found simple and effective
will be discussed here. These are lead and lag compensation. Lead compen-
sation acts mainly to speed up a response by lowering rise time and decreasing
the transient overshoot. Lag compensation is usually used to improve steady-
state accuracy of the system.
The trouble with choosing D(s) based on only a zero is that the physical re-
alization would contain a differentiator that would greatly amplify the high
frequency noise present from the sensor signal. To remedy this we simply add
a high frequency pole, perhaps at s = −20 to give
s+2
D(s) = K
s + 20
5.6. DYNAMIC COMPENSATION 113
1
Figure 5.25: Root locus for G(s) = without compensation (solid line), and
s(s + 1)
with compensation D(s) = s + 2 (dashed lines).
The resulting transfer function is thus lead compensation. The root locus with
such compensator is shown in Figure 5.26.
1 s+2
Figure 5.26: Root locus for G(s) = with lead compensation D(s) = .
s(s + 1) s + 20
To see the effect of the pole on the compensation consider moving the pole fur-
ther to the right at s = −10, i.e, nearer to the zero. The root locus is shown
in Figure 5.27. Notice the effect of moving the pole nearer to the zero, we are
reducing the effect of the zero we placed earlier. In fact, we are moving back
114 CHAPTER 5. ROOT-LOCUS ANALYSIS AND DESIGN
1 s+2
Figure 5.27: Root locus for G(s) = with lead compensation D(s) = .
s(s + 1) s + 10
to the uncompensated shape. If we move the pole too far to the left, the mag-
nification of noise at the output of D(s) is too great, since the differentiator
will dominate the compensator. Therefore, the choice of pole location is a com-
promise between the conflicting effects of noise suppression and compensation
effectivness.
The compensator pole position can now be determined by the angle criterion as
shown in Figure 5.28
α − (θ1 + θ2 + β) = ±180◦
90◦ − (135◦ + 116◦ ) ± 180◦ = β
5.6. DYNAMIC COMPENSATION 115
s0 = −2 + j2
p −2 −1
Design a lead compensator for the system given by the transfer function Example 5.8
1
G(s) =
s(s + 1)
that will provide a closed-loop damping ratio ζ = 0.5 and natural frequency
ωn > 7 rad/sec.
116 CHAPTER 5. ROOT-LOCUS ANALYSIS AND DESIGN
p −3.5
at s = 0 to raise an error constant and that is nearly unity (no effect) at the
higher frequencies.
Consider a system whose open loop transfer function is given by Example 5.9
K
G(s) =
s(s + 2)
Design a lag compensator so that the dominant poles of the closed loop system
are located at s = −1 ± j and the steady state error to a unit ramp input is less
than 0.2.
Solution For the specification that the steady state error of the system
must not exceed 0.2, we have
Kv = lim sKD(s)G(s)
s→0
(s + z) 1
= lim sK
s→0 (s + p) s(s + 2)
Kz
=
2p
2p
We require the steady state error to be less than 0.2, i.e., < 0.2. Let us
Kz
choose p = 0.01, therefore we have Kz = 0.1. We know that the closed loop
poles s = −1 ± j lie on the root locus, hence
s(s + 2)(s + 0.01)
K=−
(s + z)
s=−1+j
Solving for K and z, we get K = 1.88 and z = 0.0532. Therefore, the lag
compensator is given by
(s + 0.532)
D(s) = 1.88
(s + 0.01)
118 CHAPTER 5. ROOT-LOCUS ANALYSIS AND DESIGN
Chapter 6
Frequency Response Analysis
Frequency response methods are among the most useful techniques available for
control system analysis and design. There is no one systematic design proce-
dure for all control problems, rather, the different techniques complement each
other. Root-locus techniques give powerful indicators for closed-loop transient
response. Unfortunately, we need accurate, hence expensive, plant models to
benefit from the root locus techniques. One of the advantages of the frequency
response methods is that the response of the system can be obtained from mea-
surements on the physical system without deriving the system transfer function.
In fact, it is possible to design a control system without the need for a transfer
function model.
r(t) = A cos ωt
Then
As
R(s) =
s2 + ω 2
and
As
Y (s) = G(s)R(s) = G(s)
(s − jω)(s + jω)
We can expand this expression into partial fractions of the form
k1 k2
Y (s) = + + F (s) (6.1)
s − jω s + jω
119
120 CHAPTER 6. FREQUENCY RESPONSE ANALYSIS
where F (s) is the collection of all terms in the partial fraction that originate
in the denominator of G(s). It is assumed the system poles are real, distinct
and are in the LHP implying that the terms in F (s) will decay to zero with
increasing time. Therefore, only the first two terms in (6.1) contribute to the
steady-state response. Using the cover-up method to find k1 and k2 we have
1
k1 = (s − jω)Y (s)s=jω = AG(jω)
2
1
k2 = (s + jω)Y (s)s=−jω = AG(−jω)
2
and k2 is seen to be the complex conjugate of k1 . For any given value of ω,
k1 and k2 are complex numbers and will find it convenient to express them in
polar form as
A
k1 = |G(jω)|ejφ
2
where |G(jω)| is the magnitude and φ = G(jω). Then y(t) = L−1 {Y (s)} and
its sinusoidal steady-state value (i.e. lim y(t)) is
t→∞
We see then that, from the complex function G(jω), we can obtain the steady-
state response for any sinusoidal input, provided that the system is stable. We
call G(jω), 0 ≤ ω ≤ ∞, the frequency response function. We usually plot G(jω)
versus ω in some form to characterize the frequency response. We illustrate two
forms by a simple example. Suppose we have a system with transfer function
1
G(s) =
s+1
6.1. FREQUENCY RESPONSE 121
1 1
G(jω) = =√ − tan−1 (ω) (6.2)
1 + jω 1 + ω2
ω G(jω)
0 1.000 0◦
10 0.100 −84.3◦
∞ 0.000 −90◦
these values are plotted in the complex plane, as shown in Figure 6.1.
Note that mathematically, the frequency response is a mapping from the s-plane
to the G(jω)-plane. The upper half of the jω-axis which is a straight line, is
mapped into the complex plane via the mapping G(jω). A second form for
displaying the frequency response is to plot the magnitude and phase of G(jω)
versus ω. These plots for the example above are shown in Figure 6.2.
symbol for frequency will become evident later. Also, we call the value ωi a
break frequency, for a reason to be explained later.
|K||1 + jω/ω3 |
|G(jω)| = (6.4)
|1 + jω/ω1 ||1 + jω/ω2 |
Note that when expressed in decibels, the reciprocal of a number differs from
its value only in sign; that is,
1
20 log a = −20 log
a
1 The unit was first defined as bel, however, this unit proved to be too large, and hence a
This term is plotted versus log ω in Figure 6.4. Note that the value
√ of the term at
the frequency ωi (called the break frequency) is equal to 20 log 2 = 3.0103. We
usually approximate this value as 3dB and say that, for a general first-oder nu-
merator term, the value of the magnitude is equal to 3dB at the break frequency.
For a first-order denominator term, the value is equal to −3dB√ at its break fre-
quency. Note that the first-order term has a value of 20 log 101 = 20.04, or
approximately 20dB, at the frequency 10ωi .
Accurate Bode diagrams are usually done using digital computers. However,
there are situations in which approximate sketches of a Bode diagram are ade-
quate. We now develop the approximations for the first-order terms. Consider
the first order term of (6.6)
v"
u 2 #
u ω
20 log t 1 +
ωi
For low frequencies the term is approximated by a straight line (the ω-axis). For
high frequencies and if ω = 10ωi the difference between the logarithmic gains is
20dB. This represents a line that has a slope of 20dB per decade of frequency.
Equating the above high-frequency and low-frequency expressions shows that
the two straight lines intersect at ω = ωi . The two terms are plotted in Figure
6.5(a). Comparing this figure with the exact curve of Figure 6.4, we see that the
exact curve approaches the straight lines asymptotically, as is shown in Figure
6.5(b). As an approximation in sketching, we quite often extend the straight
lines to the intersection at ω = ωi and use this straight line approximation in-
stead of the exact curve. The frequency ωi is called the break frequency because
of the break in the slope at that frequency, as shown in Figure 6.5(b).
20 log |K|
this term does not vary with frequency. The two possible cases are shown in
Figure 6.6. If |K| is greater than unity, the magnitude is positive; if |K| is less
than unity, the magnitude is negative. In either case, the magnitude plot is a
straight line with a slope of zero.
126 CHAPTER 6. FREQUENCY RESPONSE ANALYSIS
For the case that the transfer function has a pole at the origin, the magnitude
of the term is given by
1
20 log = −20 log ω
jω
and the curve is the negative of that for a zero at the origin. Thus the curve is
a straight line with a slope of −20dB/decade that intersects the log ω axis at
ω = 1. This curve is shown in Figure 6.7(b). For the case of N th-order zeros at
the origin, the magnitude is
20 log |(jω)N | = 20 log ω N = 20N log ω
Thus the curve is still a straight line that intersects the ω-axis at ω = 1, but
the slope is now 20N dB per decade. For example if we have two zeros at the
origin the slope is 40dB/decade and -40dB/decade if two poles are at s = 0.
6.2. BODE DIAGRAMS 127
This straight line approximation is shown in Figure 6.8(a) for a zero and in
Figure 6.8(b) for a pole. Note that the terms have been normalized to have
a dc gain of unity, or 0 dB. This is convenient otherwise each term will have
a different low-frequency gain, and the Bode diagram is somewhat difficult to
plot.
Suppose that first-order term is repeated, that is, suppose that we have an
N th-order term of the form (1 + s/ωi )N . The magnitude term is then given by
" 2 #N/2 (
ω 0 ω ωi
20 log 1 + ≈ (6.7)
ωi 20N log ω/ωi ω ωi
The straight line approximation for this term is shown in Figure 6.9 for the case
of a numerator term. It is seen that for ω > ωi , the line has a slope of 20N .
For a given denominator term, the slope is −20N .
Example 6.2 Plot the Bode diagram for the system with the transfer function
10(s + 1)
G(s) =
(s + 10)
Solution First we convert the function to the form of (6.3)
(1 + s)
G(s) =
(1 + s/10)
The break frequency of the numerator is ω = 1, and the freak frequency of the
denominator is ω = 10. The numerator term, the denominator term, and the
total magnitude (which, from (6.5), is the sum of the two terms) are shown in
Figure 6.10.
The bode diagram has three terms. The first term is the constant gain, which
adds a term of value 20 log 2 = 6dB at all frequencies. The second term is the
zero term with break frequency at ω = 1, and the third term is the second-order
pole at ω = 10. The three terms are plotted in Figure 6.11.
For a real zero of the transfer function, with the zero not at the origin, the term
is given by
s 2
s jω ω
1+ = 1 + = 1 + Θ(ω)
ωi s=jω
ωi ωi
where
−1 ω
Θ(ω) = tan
ωi
Figure 6.13 shows the phase Θ plotted for various values of the ratio ω/ωi . The
exact curve is approximatted with the straight line shown in Figure 6.13. The
straight line approximation for the phase characteristic breaks from 0◦ at the
frequency 0.1ωi and breaks back to the constant value of 90◦ at 10ωi . Note that
the phase characteristic for a pole is the negative of that for a zero, since, for a
pole,
1 1 1
= =p Θ(ω)
1 + s/ωi s=jω
1 + jω/ωi 1 + (ω/ωi )2
where
−1 ω
Θ(ω) = − tan
ωi
(1 + s)
G(s) =
(1 + s/10)
6.2. BODE DIAGRAMS 131
As a second example illustrating the phase characteristic of the Bode diagram, Example 6.6
consider the system of Example 6.3, with the transfer function
5(1 + s/3)
G(s) =
s(1 + s/12)(1 + s/50)
The phase characteristics of the various terms, along with the total phase char-
acteristic of the system, are given in Figure 6.15.
Example 6.7 As a final example, the complete Bode diagram will be constructed for the
transfer function
(1 − s)
G(s) =
(1 + s/10)
For the zero at s = 1,
p
1 − jω = (1 + ω 2 ) Θ(ω) Θ(ω) = tan−1 (−ω)
lim G(jω) = 1 0◦
ω→0
−jω
lim G(jω) = = −10 = 10 −180◦
ω→∞ jω/10
by straight lines.
Consider first the case that ζ = 1. For this case, (6.9) has two real equal
zeros
2 2
s s s
1 + 2ζ + = 1+ (6.10)
ωn ωn
ζ=1 ωn
Since the zeros are real, this case is covered by the methods given in the preced-
ing sections. The straight-line approximations for this case are given in Figure
6.17, along with the exact curves. For cases in which 0 < ζ < 1, the asymptotic
approximations to the frequency response curves are not accurate and the errors
can be large for low values of ζ. This is because the magnitude and phase of
(6.9) depend on both the break frequency and the damping ratio ζ. Noting that
the exact magnitude of (6.9) in dB is
2 s 2 2
jω jω ω2 ω
20 log 1 + 2ζ + = 20 log 1 − + 2ζ
ωn ωn ωn2 ωn
134 CHAPTER 6. FREQUENCY RESPONSE ANALYSIS
Figure 6.18 illustrates some exact curves for several values of ζ between zero and
unity for complex zeros. Once again, the curves for complex poles are obtained
by inverting these curves.
For the case that ζ < 0.3, the straight line approximations are very inaccurate
and are seldom used. Instead exact curves such as in Figure 6.18 are used. An
example is now given to illustrate complex terms ina Bode diagram.
the exact Bode diagram are given in Figure 6.19. The maximum error in the
magnitude diagram for the straight line approximation is seen to be approxi-
mately 8 dB. Note also the very large errors in the straight line approximation
for the phase.
the entire frequency range in a single plot. Table 6.2 shows examples of Nyquist
plots of simple transfer functions.
G(s)
H(s) =
1 + G(s)
1 + G(s) = 0
138 CHAPTER 6. FREQUENCY RESPONSE ANALYSIS
G(s) n(s)/d(s)
H(s) = =
1 + G(s) 1 + n(s)/d(s)
n(s)
=
d(s) + n(s)
n(s)
F (s) = 1 + G(s) = 1 + =0
d(s)
d(s) + n(s)
=⇒ F (s) = =0
d(s)
That is
1 The poles of F (s) are the open loop poles (i.e., poles of G(s)).
2 The zeros of F (s) are the closed loop poles (i.e., poles of H(s)).
s − 0.5
F (s) =
s(s − 1)(s + 4)
Note that the contour in the s-plane, where F was evaluated, was traversed
in the counterclockwise direction, and enclosed the circular region in the s-
plane. Further, the contour Γ generated by the evaluation of F along Ω evolves
in the clockwise direction. Also note the contour Γ encircles the origin of the
F -plane.
(a) (b)
Figure 6.22: (a) Curve Ω in the s-plane and (b) resulting curve Γ in the F -plane.
140 CHAPTER 6. FREQUENCY RESPONSE ANALYSIS
θ F (2 θ)
0◦ 0.125 0◦
Based on the above observations, one can ask is there a relationship between
the number of poles and zeros encircled by Ω in the s-plane and the number
and direction of encirclements of the origin in the F -plane. In the above map-
ping, the counterclockwise encirclement of two poles and one zero resulted in
one clockwise encirclement of the origin.
The relationship between the contours in the two complex planes is given by
Cauchy’s theorem (known as Cauchy’s principle of argument) twhich states
(given here without proof) ”for a given contour in the s-plane that encircles
P poles and Z zeros of the function F (s) in a clockwise direction, the resulting
contour in the F-plane encircles the origin a total of N times in a clockwise
direction, where N = Z − P ”. This theorem explains the mapping in Figure
6.22, since Z = 1 and P = 2, hence, N = −1. Therefore, the contour Γ encircles
the origin once and the negative sign implies opposite direction to the contour Ω.
We now develop the Nyquist criterion. Suppose that we let the mapping of
F (s) be the characteristic polynomial of the closed-loop system of Figure 6.21;
that is;
F (s) = 1 + G(s)
Furthermore, let the curve Ω be as shown in Figure 6.23(a). This curve, which
is composed of the imaginary axis and an arc of in finite radius, completely
encircles the right half of the s-plane. Then, in Cauchy’s principle of argument,
Z is the number of zeros of the system characteristic polynomial in the right
half of the s-plane. Also recall that Z is the number of poles of the closed-loop
6.3. NYQUIST PLOTS 141
system in the RHP. Therefore, Z must be zero for the system to be stable. P
is the number of poles of the characteristic polynomial in the right half of the
s-plane and thus is the number of poles of the open loop function G(s) in the
right half plane, since the poles of 1 + G(s) are also those of G(s).
The curve in Figure 6.23(a) is called the Nyquist path, and a typical mapping
is shown in Figure 6.23(b). The mapping encircles the origin two times in the
clockwise direction, and from Cauchy’s principle
N =2=Z −P
or
Z =2+P
Since P is the number of poles of a function inside the Nyquist path, it cannot
be a negative number. This in this example, Z is greater than or equal to 2,
and the closed loop system is unstable.
142 CHAPTER 6. FREQUENCY RESPONSE ANALYSIS
Then
5
G(jω) =
(1 + jω)3
An evaluation of this function is given in Table 6.4 for certain values of ω, and a
plot of theses values is shown in Figure 6.24 The dc gain, G(0), is equal to 5 and
is shown as part I. The solid curve, part II, is obtained directly by plotting the
values of Table 6.4. However, note as ω is increased the magnitude of each factor
in the denominator has an increasing magnitude. Therefore |G(jω)| decreases
from 5 to 0.
ω G(jω)
0 5.00 0◦
20 0.0006 −261.3◦
6.3. NYQUIST PLOTS 143