Sunteți pe pagina 1din 22

1

Compensator Design to Improve Transient


Performance Using Root Locus
Prof. Guy Beale
Electrical and Computer Engineering Department
George Mason University
Fairfax, Virginia

C ONTENTS
I

INTRODUCTION

II

DESIGN PROCEDURE
II-A
Compensator Structure . . . . . . . . . . . . . . . . . . . . . . . .
II-B
Outline of the Procedure . . . . . . . . . . . . . . . . . . . . . . .
II-C
System Type N . . . . . . . . . . . . . . . . . . . . . . . . . . . .
II-D
Selecting a Dominant Closed-Loop Pole Location . . . . . . . . . .
II-E
Determining the Compensators Parameters . . . . . . . . . . . . .
II-E.1
Plant Phase Shift at s1 . . . . . . . . . . . . . . . . . .
II-E.2
Compensator Phase Shift at s1 . . . . . . . . . . . . . .
II-E.3
Placing the Compensator Zero . . . . . . . . . . . . . .
II-E.4
Placing the Compensator Pole . . . . . . . . . . . . . .
II-E.5
Determining the Compensator Gain . . . . . . . . . . .
II-E.6
Compensator Phase Shift Revisited . . . . . . . . . . . .
II-E.7
Simultaneous Placement of Compensator Pole and Zero
II-E.8
Multi-Stage Compensation . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

2
2
3
4
4
6
6
7
8
9
10
11
13
13

Design Example
III-A
Phase Lead Example . . . . . . . . . . . . . . . . . . .
III-A.1
Given System and Specifications . . . . . . .
III-A.2
Selection of the Dominant Closed-Loop Pole
III-A.3
Designing the Compensator . . . . . . . . . .
III-B
Phase Lag Example . . . . . . . . . . . . . . . . . . . .
III-B.1
Given System and Specifications . . . . . . .
III-B.2
Selection of the Dominant Closed-Loop Pole
III-B.3
Designing the Compensator . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

16
16
16
16
16
18
18
19
19

III

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

References

1
2
3
4
5
6
7
8
9
10
11
12

L IST OF F IGURES
Allowable region for s1 in order to satisfy specifications on overshoot, settling time, and frequency of
Calculating the phase shift of G(s) at the chosen point s = s1 . . . . . . . . . . . . . . . . . . . . . .
One possible solution for locating the compensator zero and pole. . . . . . . . . . . . . . . . . . . . .
Comparison of closed-loop step responses for 5 compensator designs with s1 = 4 + j5.4575. . . . .
Root locus plot for four compensator designs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Comparison of lag and lead compensation for a particular system. . . . . . . . . . . . . . . . . . . . .
Selecting the compensator zero and pole to maximize = zc /pc . . . . . . . . . . . . . . . . . . . . .
Step response of the uncompensated system for the phase lead design example. . . . . . . . . . . . .
Root locus of the uncompensated system and the desired closed-loop pole s1 = 0.25 + j0.488. . . .
Lead compensated root locus and step response for the design example. . . . . . . . . . . . . . . . .
Comparison of root locus plots with two lag compensator designs. . . . . . . . . . . . . . . . . . . .
Comparison of closed-loop step responses for two lag compensators. . . . . . . . . . . . . . . . . . .

22

oscillation.
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .

6
7
9
11
12
14
15
17
18
19
21
22

These notes are lecture notes prepared by Prof. Guy Beale for presentation in ECE 421, Classical Systems and Control Theory, in the Electrical and
Computer Engineering Department, George Mason University, Fairfax, VA. Additional notes can be found at: http://teal.gmu.edu/~gbeale/examples.html.

I. INTRODUCTION
In my approach to root locus design, the purpose of compensator design using root locus methods generally is to establish a
specified point in the s-plane, s = s1 , as a closed-loop pole. The assumption is that time-domain transient specifications, such
as settling time and overshoot, will be satisfied if s1 is a dominant closed-loop pole. In the simplest case, s1 is already on
the root locus of the uncompensated system. The compensator is then just a gain Kc that is chosen to satisfy the magnitude
criterion at the point s1 .
More often, the point s1 is not on the uncompensated root locus, so the compensator must add enough phase shift at the
point s1 to satisfy the phase angle criterion so that the compensated root locus does pass through s1 . This is done by choice
of the compensators poles and zeros. The compensator gain is then chosen to satisfy the magnitude criterion at the point s1 .
In many cases, the speed of response and/or the damping of the uncompensated system must be increased in order to satisfy
the specifications. This requires moving the dominant branches of the root locus to the left. A phase lead compensator (providing
positive phase shift at s1 ) is used for this purpose. If the branches need to be moved to the right, a phase lag compensator
(providing negative phase shift at s1 ) is used. The design techniques are identical for the two types of compensator; the roles of
the compensators poles and zeros are just reversed. Because of this similarity in the design methods, phase lead compensation
will be discussed in detail here. An example of each type of compensation will be given after the general design procedure is
described.
Conceptually, the design procedure presented here is graphical in nature. The process of locating the compensators poles
and zeros to satisfy the phase requirements can be visualized from the trigonometric relationships that must be satisfied at the
desired dominant closed-loop pole. The computations can be easily done by calculator. If data arrays representing the numerator
and denominator polynomials of the open-loop system are available, then the procedure can be done using a software package
such as MATLAB, and in many cases it can be automated. The examples and plots presented here are all done in MATLAB,
and the various measurements that are presented in the examples are obtained from the arrays storing the appropriate variables.
The primary references for the procedures described in these notes are [1][3]. Other references that contain similar material
are [4][11].
II. DESIGN PROCEDURE
A. Compensator Structure
The basic phase lead or phase lag compensator consists of a gain, one real pole, and one real zero. Based on the usual
electronic implementation of the compensator [3], the circuit for a lead or lag compensator is the series combination of two
inverting operational amplifiers. The first amplifier has an input impedance that is the parallel combination of resistor R1 and
capacitor C1 and a feedback impedance that is the parallel combination of resistor R2 and capacitor C2 . The second amplifier
has input and feedback resistors R3 and R4 , respectively.
Assuming that the op amps are ideal, the transfer function for this circuit is

s
+1
Kx
Vout (s)
K ( s + 1)
Kc (s zc )
zc
= x
Gc (s) =
=
=
(1)
s
Vin (s)
(s pc )
( s + 1)
+1
pc
R4 R2 (sR1 C1 + 1)

=
R3 R1 (sR2 C2 + 1)
R4 R2 R1 C1 (s + 1/R1 C1 )

=
R3 R1 R2 C2 (s + 1/R2 C2 )
R4 C1 (s + 1/R1 C1 )

=
R3 C2 (s + 1/R2 C2 )
The zero and pole of the compensator are located at s = zc at s = pc , respectively. Therefore, zc and pc are negative if they
are located in the left-half of the s-plane and positive if they are in the right-half plane. The following relationships1 can be
obtained by inspecting Eq. (1).
R4 R2
zc
R2 C2
,
= R1 C1 ,
= R2 C2 ,
=
=
(2)
Kx =
R3 R1
pc
R1 C1
Kx
R4 C1
zc = 1/R1 C1 ,
pc = 1/R2 C2 ,
Kc =
=

R3 C2
The zero zc is to the right of the pole pc for a phase lead compensator, and it is to the left of the pole for a phase lag
compensator. This is true whether the pole and zero are in the left-half plane (the usual case) or in the right-half plane.
1 In my descriptions of the design of compensators using Bode plots (Phase Lag Compensator Design Using Bode Plots, Phase Lead Compensator Design
Using Bode Plots), a slightly different definition for the compensator transfer function is used, although it refers to the same electronic circuit shown in [3].

Assuming that zc < 0 and pc < 0 (both in the left-half plane), then < 1 for a lead compensator, and > 1 for a lag
compensator.
At any point s = s1 , the compensator provides a magnitude
q
2
|Kc | [Re (s1 ) zc ] + Im2 (s1 )
|Kc | |s1 zc |
q
(3)
|Gc (s1 )| =
=
|s1 pc |
2
[Re (s1 ) pc ] + Im2 (s1 )

and a phase angle

Im (s1 )
Im (s1 )
1
]Gc (s1 ) = ]Kc + ] (s1 zc ) ] (s1 pc ) = ]Kc + tan
tan
(4)
Re (s1 ) zc
Re (s1 ) pc
Only positive values of compensator gain will be discussed in these notes, so with Kc > 0, the magnitude and phase of the
gain are |Kc | = Kc and ]Kc = 0 . In applications where Kc < 0, the we have |Kc | = Kc and ]Kc = 180 .
1

B. Outline of the Procedure


The following steps outline the procedure that will be used to design either a lead or a lag compensator using root locus
methods in order to satisfy transient performance specifications, such as settling time and percent overshoot. Compensator
design to satisfy steady-state error requirements is discussed in separate notes, Compensator Design to Improve Steady-State
Error Using Root Locus.
1) Determine if the System Type N needs to be increased in order to satisfy the steady-state error specification, and if
necessary, augment the plant with the required number of poles at s = 0. This should be done at this point in the design,
even though numerically satisfying the steady-state error specification is done later. If the System Type is not taken care
of now, the remainder of the design procedure may be ineffective when the System Type is changed at the end of the
design.
2) Choose a point in the s-plane to be the location for a dominant closed-loop pole. The selection of this point is based on
the transient performance specifications and should produce a closed-loop system that will satisfy those specifications.
3) Design the compensator:
a) Compute the phase shift of the plant (including any additional poles at s = 0 needed to satisfy the steady-state error
specification) at the chosen point s = s1 . If the phase shift is an odd integer multiple of 180 (180 mod 360 ), then
the selected point is on the uncompensated systems root locus for positive gain. If the phase shift is an even integer
multiple of 180 (0 mod 360 ), then the selected point is on the uncompensated systems root locus for negative
gain. In either case, the only compensation needed for the transient performance specifications is the proper gain,
and the procedure can jump to step 3(e).
b) Assuming that the selected point s1 is not already on the root locus (for either positive or negative gain), compute
the amount of phase shift that the compensator must provide at s = s1 in order to make that point lie on the root
locus. The compensators phase shift (usually) will be the shortest distance from the plants phase shift at s1 to
an odd integer multiple of 180 for positive gain (the most usual case) or to an even integer multiple of 180 for
negative gain. It should be noted that having the compensator provide this amount of phase shift only guarantees
that the point s1 will lie on the compensated root locus; it does not guarantee that the closed-loop system will be
stable when s1 is a closed-loop pole. Additional factors must be taken into account for that.
c) Select the location for either the compensator zero zc or pole pc . If the compensator phase shift at s1 computed
in the previous step is positive, phase lead compensation is required. This places the compensator zero to the right
of the pole. If the compensators phase shift is negative, then phase lag compensation is needed, and the zero is to
the left of the pole. One of these factors (zero or pole) can be placed with some freedom; once it is placed, the
location of the other factor is fixed.
d) Compute the horizontal distance from the point s = s1 to the location of the compensator factor not already placed
(pole or zero), and place that factor at that location. This pole/zero pair in series with the plant makes the root
locus pass through the point s1 .
e) Compute the compensator gain Kc . The value of the gain is chosen to satisfy the magnitude criterion at s1
(|Gc (s1 ) Gp (s1 )| = 1) in order for s1 to be a closed-loop pole.
4) If necessary, choose appropriate resistor and capacitor values to implement the compensator design.
To illustrate the design procedure, the following system model and specifications will be used:
8
Gp (s) =
s+4
steady-state error specification for a unit ramp input is ess_specif ied = 0.1;
step response settling time specification is Tsspecif ied = 1 second;
step response overshoot specification P Ospecif ied 10%.

(5)

C. System Type N
The first step in the design of the compensator will be to determine if the plant Gp (s) has the correct System Type to
satisfy the steady-state error specification. Defining the number of open-loop poles of a system that are located at s = 0 to
be the System Type N , and restricting the reference input signal to having Laplace transforms of the form R(s) = A/sq , the
steady-state error and error constant are (assuming that the closed-loop system is bounded-input, bounded-output stable)
N +1q
As
ess = lim N
(6)
s0 s + Kx
where

Kx = lim sN G(s)
s0

(7)

For N = 0, the steady-state error for a step input (q = 1) is ess = A/ (1 + Kx ). For N = 0 and q > 1, the steady-state
error is infinitely large. For N > 0, the steady-state error is ess = A/Kx for the input type that has q = N + 1. If q < N + 1,
the steady-state error is 0, and if q > N + 1, the steady-state error is infinite. Therefore, for a system to have a non-zero, finite
steady-state error for a specified reference input, the System Type must satisfy N = Nreq = q 1. This is the total number of
open-loop poles at the origin needed for the compensated system to satisfy the steady-state error specification. If the System
Type of the plant Gp (s) is Nsys , then the compensator must have (Nreq Nsys ) poles at the origin. These poles would be
included with the plant model during the design of the rest of the compensator, and then they would be implemented as part
of the compensator after the design is complete. The system that will be evaluated during the rest of the design process will
be
1
G(s) = (N N ) Gp (s)
(8)
req
sys
s
Example 1:
The plant transfer function in (5) has no poles at the origin, so it has System Type Nsys = 0. The steady-state error
specification is for a unit ramp input which has a Laplace Transform R(s) = 1/s2 , so for this input q = 2. Therefore, in order
for the steady-state error specification to be satisfied, the total number of poles at the origin must be Nreq = q 1 = 2 1 = 1.
Thus, the compensator must provide Nreq Nsys = 1 0 = 1 pole at the origin. The system G(s) corresponding to Eq. (8)
for this example is G(s) = 8/ [s (s + 4)].

Note that satisfying the numerical value of the steady-state error is not done at this point in the design. That will be done
as a separate task.
D. Selecting a Dominant Closed-Loop Pole Location
The real engineering design takes place in this step. This is the mapping of the performance specifications that must be
satisfied into the design parameter that influences the rest of the steps in the design. For other than very simple systems, sound
engineering judgement is needed in order for this mapping to yield a compensator design that actually allows the specifications
to be satisfied.
In many cases, the dominant closed-loop poles are chosen to be a complex conjugate pair, so the design point s = s1 is a
complex number with negative real part and positive imaginary part. The starting point in choosing s1 generally is to make use
of the relationships between time-domain characteristics and closed-loop pole locations that exist for the standard second-order
system, shown in (9).
G(s) =

2n
,
s (s + 2 n )

TCL (s) =

2n
s2 + 2 n s + 2n

(9)

where is the dimensionless damping ratio, and n is the undamped natural frequency (rad/ sec). For 0 < < 1, the closedloop system is underdamped, the closed-loop poles are complex conjugates, and the step response exhibits overshoot and
damped sinusoidal transient behavior. The closed-loop poles are at
q
(10)
p1 , p2 = n j n 1 2

In this case, specifications on the amount of overshoot and the settling time for the step response are natural specifications.
These types of specifications will be used in the design procedure presented in these notes. A steady-state error specification
on the ramp response is also appropriate; that will be discussed in separate notes.
Percent overshoot (P O) in the step response of the standard second-order system is only a function of the damping ratio.
The value of overshoot (%) and the value of are related by
P O

h
i

ln
/ 1 2
100
PO = e
100%,
=q
(11)
O
2 + ln2 P100

Points in the s-plane that correspond to a constant value of lie on a radial line from the origin making an angle with respect
to the negative real axis, with = cos1 (). Therefore, a specification of percent overshoot establishes a fixed relationship
between the imaginary part of the dominant closed-loop pole and the real part of the pole. This relationship
is given

by tan(),
which is the slope of the radial line corresponding to the value of , so that Im [s1 ] = Re [s1 ] tan cos1 () .
Defining settling time Ts as the time required for the step response to get within and stay within a fixed percentage of the
final value, the settling time for the standard second-order system is a function of the product of damping ratio and undamped
natural frequency. In these notes, the band about the final value used to define Ts is 2%, and the settling time is
Ts =

4
n

(12)

Comparing (12) and (10) shows that settling time is related to the real part of the closed-loop poles, so a settling time
specification corresponds to the dominant closed-loop poles lying on a vertical line in the s-plane located at
4
(13)
Ts
If the settling time specification is an upper bound, then the closed-loop poles must lie on or to the left of the vertical line
given in (13).
p
The damped natural frequency is defined to be d = n 1 2 , which is seen in (10) to be the imaginary part of the
closed-loop poles. This parameter is also the frequency of oscillation in the transient part of the step response. Therefore, a
specification on the frequency of oscillation imposes a restriction on the imaginary part of the dominant closed-loop poles,
which corresponds to a horizontal line in the s-plane.
If two or more specifications are imposed on the system, then there are multiple constraints on the location of the dominant
closed-loop pole s1 . An acceptable location for that pole must satisfy each of the constraints. Therefore, s1 must lie in the
intersection of the regions defined by the various specifications.
Re [s1 ] =

Example 2:
The following specifications are to be imposed on a systems closed-loop step response: (1) 5% P O 25%, (2) 0.5 s
Ts 2 s, (3) d 12.6 r/s. Assuming that the equations for the standard second-order system will hold for the actual system,
then the following constraints are imposed on the location of s1 . From (11), the lower bound of 5% overshoot corresponds to
a damping ratio = 0.6901 and an angle = 46.4 . The point s = s1 must lie on or above this radial line. The upper bound
of 25% overshoot corresponds to a damping ratio = 0.4037 and an angle = 66.2 . The point s = s1 must lie on or below
this radial line. From (13), the lower bound of Ts = 0.5 s corresponds to real part of Re [s1 ] = 8. The point s1 must lie on
or to the right of this vertical line. The upper bound of Ts = 2 s corresponds to real part of Re [s1 ] = 2. The point s1 must
lie on or to the left of this line. The upper bound of d = 12.6 r/s corresponds to a horizontal line at j12.6. The closed-loop
pole must lie on or below this line. The dominant closed-loop pole s1 must lie in the intersection of these 5 constraint curves.
The dotted region in Figure 1 is the set of acceptable locations for s1 for this example.

Example 3:
In Example 1 the original Gp (s) augmented with one pole at the origin to satisfy the steady-state error requirement is
G(s) = 8/[s(s + 4)]. The transient specifications are Tsspecif ied = 1 second and P O 10%. The damping ratio associated
with 10% overshoot is = 0.5912. The corresponding angle and slope of the radial line from the origin through s1 are
= 53.76 and 1.3644, respectively. From (13), the real part of s1 is s = 4. Combining these two requirements places the
dominant closed-loop pole at s1 = 4 + j5.4575.

The relationships between dominant pole locations and step response characteristics presented above are only for the standard
second-order system of (9). These relationships may provide good starting points for the selection of s1 to satisfy specifications,
but they can only be used as general guidelines for more complex system models. Even if the given system Gp (s) is modeled
by (9), any compensation in series with Gp (s), other than a pure gain, will result in a more complex model. The overshoot
equation (11) generally is not a good prediction of what the actual overshoot will be in a higher-order system or in a system
with zeros. The actual overshoot may be less than predicted by (11), but it will more likely be larger. In choosing a value for s1
to satisfy a percent overshoot specification, the recommended approach is to be very conservativethat is, use a significantly
smaller value of overshoot than the specified value when computing the effective damping ratio from (11).
The prediction of the real part of s1 computed from (13) to satisfy a settling time specification is generally fairly accurate.
This is due to the fact that the decay of the transient response is controlled by the real part of the dominant closed-loop poles.
This is how the expression in (13) was derived and accounts for its accuracy in predicting the settling time of higher-order
systems.
In many actual design problems with specifications on overshoot, settling time, frequency of oscillation, etc., the proper
choice of location for the dominant closed-loop pole (or complex conjugate pair) will be achieved only through iteration.
Engineering design is an iterative process, and the selection of s1 is a major step in the design of the compensator that will
allow the specifications to be satisfied.

Locating ClosedLoop Poles in the splane to Satisfy Specifications


20
PO

POmin = 5%

max

max = 0.6901

15

min

= 25%

min = 0.4037

= 46.4

max

= 66.2

= 12.6 r/s

max

Imag Axis

10

0
T

= 0.5 s

min

max

= 8
n

5
20

15

10

=2s

Ts

= 2
n

Real Axis
Fig. 1.

Allowable region for s1 in order to satisfy specifications on overshoot, settling time, and frequency of oscillation.

E. Determining the Compensators Parameters


Once the point s = s1 is chosen to be a dominant closed-loop pole, the design of the compensator is an exercise in
trigonometry. The pole(s) and zero(s) of the compensator are chosen to satisfy the phase angle criterion at s1 so that the total
phase shift of the series combination of the plant and compensator at s1 is an odd integer multiple of 180 if K > 0 or an
even integer multiple of 180 if K < 0. This makes the root locus pass through the point s1 . Once the pole(s) and zero(s) are
placed, the gain of the compensator is selected to satisfy the magnitude criterion. This makes the point s = s1 be a closed-loop
pole. However, as previously mentioned, making s = s1 be a closed-loop pole does not guarantee that the closed-loop system
is stable. There may still be closed-loop poles in the right-half plane. Thus, there is more to choosing the locations of the
compensators pole(s) and zero(s) than just satisfying the phase angle criterion.
1) Plant Phase Shift at s1 : The starting point in the compensator design is to determine the phase shift of the augmented
plant of Eq. (8) at the point s1 . The phase shift of G(s) at s1 is
]G (s1 ) = ]K + ]N (s1 ) ]D (s1 )
m
n
X
X
= ]K +
] (s1 zi )
] (s1 pi )
i=1

= ]K +

m
X
i=1

tan1

i=1

n
Im (s1 ) Im (zi )
1 Im (s1 ) Im (pi )
tan

Re (s1 ) Re (zi )
Re (s1 ) Re (pi )
i=1

where the zi and pi are the open-loop zeros and poles, respectively. We will assume that K > 0, so ]K = 0.

(14)

G(s) = 8/[s(s+4)], s1 = 4 + j5.4575


6
s1

Imag Axis

p2

p1

6
10

Fig. 2.

4
Real Axis

Calculating the phase shift of G(s) at the chosen point s = s1 .

Care must be used when evaluating the tan1 function when using the normal atan function in MATLAB or on calculators.
If the denominator of the tan1 function is negative, the atan function will return the incorrect angle. To obtain the correct
angle in this case, 180 ( rad) must be added to the angle returned by the atan function. The function atan2 returns the
correct value since it takes two input arguments; its syntax is = atan2 (Im, Re).
Example 4:
From Example 1, the augmented plant model is G(s) = 8/ [s (s + 4)], so there are no zeros zi and the open-loop poles are
p1 = 0 and p2 = 4. We will assume that the desired dominant closed-loop pole is at s = s1 = 4 + j5.4575, based on
Example 3. The phase angle of G(s) at s1 is: ]G (s1 ) = ]8 ]s1 ] (s1 + 4) = 0 tan1 [(5.4575 0) / (4 0)]
tan1 [(5.4575 0) / (4 + 4)]. ]G (s1 ) = tan1 [5.4575/ (4)] tan1 (5.4575/0) = (53.76 + 180 ) 90 =
216.24 = 143.76 . Both forms for the phase shift, ]G (s1 ) = 216.24 or ]G (s1 ) = 143.76 , are acceptable ways
to represent the angle. Figure 2 shows the relationships between the open-loop poles and the point s1 . The angles that are
computed are the angles of the vectors drawn from the poles to the point s1 measured counter-clockwise from the positive
real axis.

Note the 180 added to the phase shift returned by the atan function in the above example for the pole p1 . This will be the
case when the point s1 is to the left of the pole or zero under consideration.
2) Compensator Phase Shift at s1 : If the point s1 is on the root locus of the uncompensated system G(s) (which is Gp (s)
augmented with any poles at the origin that were needed to produce the correct System Type), then ]G (s1 ) is an odd integer
multiple of 180 (for K > 0). In this case ]Gc (s1 ) = 0 , and the only (additional) compensation that is needed in order to
make s = s1 a closed-loop pole is a gain Kc . That would be chosen to satisfy the magnitude criterion |Kc G (s1 )| = 1.

Most often, s1 will not be on the original root locus, so the compensator must also provide a phase shift at s1 so that the
total phase shift of the plantcompensator combination is an odd integer multiple of 180 at that point. This is done by choice
of the compensators pole and zero. Satisfying the phase angle criterion must be done first. After that, the value of Kc is
computed to satisfy the magnitude criterion.
Given the value ]G (s1 ) from the previous step, the required compensator angle ]Gc (s1 ) can be easily computed. Generally,
the compensator phase shift at s1 is chosen to be the angle with the smallest absolute value such that ]G (s1 ) + ]Gc (s1 )
is an odd integer multiple of 180 . Thus, if ]G (s1 ) = 150 or ]G (s1 ) = 210 , then ]Gc (s1 ) = +30 will place s1 on
the root locus, and the compensator is phase lead. Likewise if ]G (s1 ) = 150 or ]G (s1 ) = 210 , then ]Gc (s1 ) = 30
will place s1 on the root locus, and the compensator is phase lag. However, in some cases it might be necessary (or desirable)
to use a particular type of compensator, such as phase lead, in order to satisfy other objectives. In a situation like this, if
]G (s1 ) = 60 for example, then the compensator phase shift at s1 would be chosen to be 180 (60 ) = 240 (phase
lead) rather than 180 (60 ) = 120 (phase lag). A more detailed example illustrating this situation is given in Section
II-E.6.
The necessary compensator phase shift to make s1 lie on the root locus is given by
]Gc (s1 ) = 180 (2l + 1) ]G (s1 )
= ]Kc + ]Nc (s1 ) ]Dc (s1 )
mc
nc
X
X
= ]Kc +
] (s1 zci )
] (s1 pci )
i=1

(15)

i=1

Assuming that Kc > 0, then ]Gc (s1 ) = ]Nc (s1 ) ]Dc (s1 ).

Example 5:
From Example 4, the phase shift of G(s) at s1 = 4 + j5.4575 is ]G (s1 ) = 216.24 = 143.76 . The phase angle of
the compensator needed to place s = s1 on the root locus (using the smallest absolute value for the compensator phase angle)
is ]Gc (s1 ) = 180 (216.24 ) = 180 143.76 = 36.24 . Since this angle is positive, the compensator is phase lead.
The pole and zero of the compensator will be placed on the real axis such that ] (s1 zc ) ] (s1 pc ) = 36.24 . Methods
of choosing appropriate locations are presented in the next section. One possible solution is to place the compensator zero at
s = 5, which provides 79.62 of phase shift at s1 , and to place the compensator pole at s = 9.78, which provides 43.38
at s1 . The difference between those angles is the required value of 36.24 . Figure 3 illustrates this solution.

3) Placing the Compensator Zero: There is some flexibility in choosing the location of the compensator zero, but not
complete freedom. Assuming that a single stage of compensation is desired, the left-most location of the zero is governed by
the compensators phase shift requirement at s1 . Because ]Gc (s1 ) = ] (s1 zc ) ] (s1 pc ), the compensator zero must
provide more phase shift at s1 than the total compensator. Therefore, the left-most allowed position of the compensator zero
is constrained by ]Gc (s1 ).
The right-most location of the zero is generally constrained by transient performance specifications, and ultimately by the
requirement for closed-loop stability. Moving the compensator zero to the right also moves a closed-loop pole to the right.
For some choices for the location of the zero, there may be a closed-loop pole closer to the j axis than the selected point
s1 . When this happens, s1 may no longer be the dominant pole in terms of settling time. If the compensator zero is in the
right-half plane, then the closed-loop system may be unstable.
Unfortunately, there is not a rule for placing the compensator zero. The amount of overshoot can vary widely depending on
its location. For systems with one open-loop pole at the origin and with one or more additional poles on the negative real axis,
a rule of thumb is to place the compensator zero at or to the left of the second real-axis open-loop pole. This rule of thumb
is intended to ensure that s = s1 is the dominant closed-loop pole, but it does not guarantee that overshoot specifications will
be satisfied.
Example 6:
For the system G(s) defined in the previous examples and the point s1 = 4 + j5.4575, the required phase shift of the
compensator was found to be ]Gc (s1 ) = 36.24 . Thus, the compensator zero must provide more than 36.24 at s1 . In order
to use a single stage of compensation, the left-most point for the zero is s = Re [s1 ] [Im [s1 ] / tan (36.4 )] = 11.45.
The right-most point for the compensator zero is just to the left of the origin. The zero cannot be placed at s = 0 since
that would cancel the open-loop pole there, and internal stability would be lost. Placing the compensator zero in the righthalf plane would also result in an unstable closed-loop system. Some possibilities for the location of the zero are zc = 9
[] (s1 zc ) = 47.51 ], zc = 7 [] (s1 zc ) = 61.20 ], zc = 5 [] (s1 zc ) = 79.62 ], zc = 3 [] (s1 zc ) = 100.4 ],
zc = 1 [] (s1 zc ) = 118.8 ]. The effects of these different choices for zc will be shown in the following examples.

The first three locations for the zero in the above example (zc = 9, zc = 7, zc = 5) satisfy the rule of thumb mentioned
earlier. For these choices, the closed-loop poles at s = s1 and its complex conjugate will be closer to the j axis than the

Providing 36.24 degrees at s1 = 4 + j5.4575


6
s1

Imag Axis

pc

zc

6
10

Fig. 3.

4
Real Axis

One possible solution for locating the compensator zero and pole.

third pole. Therefore, the settling time of the step response will be governed by the choice of s1 , as desired. The remaining
two choices for the compensator zero (zc = 3, zc = 1) do not satisfy the rule of thumb. The poles at s1 will no longer be
closest to the j axis, and so will not be controlling the settling time. This does not mean that the settling time specification
will be violated if the rule of thumb is not satisfied, but as the zero is moved farther and farther to the right, at some point,
the settling time specification will be violated.
4) Placing the Compensator Pole: Once the location for the compensator zero has been selected, there is no more freedom in
the design; the compensator pole and gain are now fixed. The location of the pole is constrained by the phase angle requirement
on the compensator at the point s1 , and once the pole location is determined, the gain is constrained by the magnitude criterion
at s1 .
Since the total phase shift of the compensator at s = s1 is the angle produced by the zero minus the angle produced by the
pole, the phase shift of the compensator pole at s1 is
] (s1 pc ) = ] (s1 zc ) ]Gc (s1 )

(16)

The value of this angle can be computed once the location of the zero is chosen. There is only one location for the compensator
pole that will produce the angle ] (s1 pc ) at s = s1 . The distance dpc from the real-axis projection of s1 to the compensator
pole can be computed easily from
dpc =

Im [s1 ]
tan [] (s1 pc )]

(17)

10

The compensator pole is located to the left of s1 if dpc > 0 and to the right of s1 if dpc < 0. The location of the compensator
pole at s = pc is given by
pc = Re [s1 ] dpc

(18)

Example 7:
From Example 5, the compensator must provide 36.24 of phase shift at s1 . In Example 6, five possible choices for the
compensator zero were presented, namely, zc = {9, 7, 5, 3, 1}. For each of these choices, the compensator pole can
be located by using Eqs. (16) (18). The angles, distances, and locations of the compensator pole are shown in the following
table for the various choices of compensator zero.
] (s1 pc )
11.27
24.96
43.38
64.14
82.56

zc
9
7
5
3
1

dpc
27.4
11.7
5.78
2.64
0.713

pc
31.4
15.7
9.78
6.64
4.713

The table shows that as the compensator zero moves to the right, the required phase shift of the pole increases, and the pole
also moves to the right.

5) Determining the Compensator Gain: Now that the pole and zero of the compensator have been selected, the root locus
plot will pass through the point s = s1 . Therefore, s1 is a potential closed-loop pole. In order for that point to actually be
a closed-loop pole, the magnitude criterion must be satisfied at s1 by the series combination of plant and compensator. The
compensator gain is used for this purpose.
The magnitude criterion states that
|Kc | |s1 zc |
|G (s1 )| = 1
|s1 pc |

(19)

if s = s1 is a closed-loop pole, with G(s) defined in (8). The compensator gain is the only parameter that is undetermined at
this point. Therefore, to make s = s1 be a closed-loop pole, the compensator gain must be
|Kc | =

|s1 pc |
|s1 zc | |G (s1 )|

(20)

Note that (20) only provides the magnitude (absolute value) of the compensator gain. If the system requires that the gain be
negative, then Kc = |Kc |.
Example 8:
For each of the combinations of compensator zero and pole from Examples 6 and 7, there is a unique value for the gain Kc .
The value of gain that places one of the closed-loop poles at s1 is computed from (20). The magnitude of the augmented plant
at s1 is |G (s1 )| = 8/ (|s1 | |s1 + 4|) = 0.2166. Using the values for zc and pc from the table in Example 7, the corresponding
gain values are given in the following table.
zc
9
7
5
3
1

|s1 pc |
27.94
12.93
7.946
6.065
5.504

|s1 zc |
7.402
6.228
5.548
5.548
6.228

Kc
17.42
9.585
6.611
5.045
4.080

For each of these combinations, the compensator transfer function, including the pole at s = 0 needed for the steady-state
error specification, is Gc (s) = Kc (s zc ) / [s (s pc )]. With the values used here, the closed-loop system is stable, and two
of the closed-loop poles are at the specified location of s1 = 4 + j5.4575 and its complex conjugate. The location of the
third closed-loop pole varies with the choice of compensator parameters. For the values used in these examples, the third pole
is at s = {27.4, 11.7, 5.78, 2.64, 0.713}. Note that for the particular structure of Gc (s)Gp (s) in these examples, the
absolute value of the third closed-loop pole is equal to the distance between the real-axis projection of s1 and the compensators
open-loop pole. The closed-loop step responses for these five compensator designs are shown in Fig. 4. The settling times for
the first four compensator designs are all approximately the same, slightly less than 1 second. The settling time for the fifth
compensator design is significantly longer and does not satisfy the specifications. Of the compensator designs that do satisfy

11

ClosedLoop Step Responses


1.4

1.2
1.1
1.02
1

Output Amplitude

0.98

0.8

0.6

0.4
zc = 9
z = 7
c

0.2

z = 5
c

z = 3
c

z = 1
c

Fig. 4.

0.2

0.4

0.6

0.8

1
Time (s)

1.2

1.4

1.6

1.8

Comparison of closed-loop step responses for 5 compensator designs with s1 = 4 + j5.4575.

the settling time requirement, only the fourth design (zc = 3) also satisfies the overshoot specification. The dashed lines in
the figure show the 2% interval for the settling time specification and the 10% overshoot specification. Figure 5 shows the
root locus plots for four of the compensator designs. The triangles shown in the plots represent the closed-loop poles. The
point s1 is a closed-loop pole in each case. It is easily seen in the figure that the third closed-loop pole moves to the right as
zc does, and when zc = 1 the third pole is far to the right of s1 . This causes the much longer settling time evident in Fig.
4.

These examples have illustrated the procedure for mapping a set of transient response specifications into a desired location
s = s1 for the dominant closed-loop pole and for designing a compensator to make that desired location actually be a closedloop pole. The examples have shown that not every compensator design that places a closed-loop pole at s1 will satisfy all of
the specifications. Closed-loop stability is not even guaranteed if the compensated system is third-order or higher. Therefore,
the design of the compensator will generally be an iterative process. Although there are general guidelines that can be followed,
a design needs to be validated through simulation to determine whether or not it is an acceptable design.
6) Compensator Phase Shift Revisited: It was mentioned in Section II-E.2 that the compensator phase shift at s1 is generally
taken to be the smallest absolute value that will yield an odd integer multiple of 180 for the plantcompensator combination.
It was also mentioned that there are instances when such a choice would be inappropriate. The following example illustrates
when closed-loop stability requirements force a change to this general procedure.
Example 9:
Consider the system G(s) = 5(s + 2)/[s2 (s + 0.4)(s + 6)] and the desired closed-loop pole s1 = 1 + j0.8. The phase
shift of G(s) at s1 is ]G (s1 ) = 19.98 . A compensator phase ]Gc1 (s1 ) = 160.02 would place s1 on the root locus for
Kc > 0. This would require a phase lag compensator. On the other hand, a compensator phase ]Gc2 (s1 ) = 199.98 would
also place s1 on the root locus for Kc > 0. This would require a phase lead compensator. Which one should be used? In
answering this question, it should be remembered that open-loop poles tend to repel branches of the root locus, and open-loop

12

z = 7

z = 5

20

15

15

10
5

Imag Axis

Imag Axis

10

10
10

15
20
20

15

10
5
Real Axis

15
10

Real Axis

z = 3

z = 1

10

8
6
4

Imag Axis

Imag Axis

2
0
2

4
6

10
8

Fig. 5.

4
2
Real Axis

Root locus plot for four compensator designs.

8
5

2
1
Real Axis

13

zeros tend to attract them. The system G(s) has two open-loop poles at the origin, the boundary of stability. Using a phase
lag compensator, with the pole to the right of the zero (closer to the j axis assuming they are in the left-half plane) would
tend to move the branches of the root locus that begin at s = 0 into the right-half plane, resulting in an unstable closed-loop
system. A lead compensator, on the other hand, would tend to draw those branches into the left-half plane. Figure 6 illustrates
this. The top two plots are the root loci for the system with phase lag and phase lead compensation. The compensator transfer
functions are Gclag (s) = 0.39 (s + 5.54)2 / (s + 1)2 and Gclead (s) = 71.9 (s + 0.75)2 / (s + 7.19)2 . It is clearly seen that
the lag compensator produces an unstable system for all K > 0. The lead compensator is able to stabilize the system. It should
be noted that s = s1 is a closed-loop pole in both designs. The bottom two plots are the closed-loop step responses. The
instability of the design with the lag compensator is clearly indicated.

7) Simultaneous Placement of Compensator Pole and Zero: As we discussed in the previous sections, there is some freedom
in placing the compensator zero and pole, as long as the total phase angle of the compensator has the correct value. That
freedom can be utilized to vary the amount of overshoot or the value of the settling time. Unfortunately, there is generally not
a clear decision procedure for selecting the location of the pole or zero.
There is a procedure that places both the pole and zero at the same time. There is no freedom of choice with this procedure,
but it does have the advantage of maximizing the value of = zc /pc for a lead compensator or minimizing for a lag
compensator. This is advantageous because it minimizes the range of resistor and capacitor values used to implement the
compensator.
We will define the phase shift of the desired closed-loop pole at s1 to be the angle of the radial line drawn from the origin
to s1 , measured in the counter-clockwise direction from the positive real axis, and denote it by ]s1 . With that definition, the
phase shifts of the compensator zero and pole at s1 are
]s1 + ]Gc (s1 )
]s1 ]Gc (s1 )
,
] (s1 pc ) =
2
2
Therefore, the distances from the real-axis projection of s1 to the zero and pole are
] (s1 zc ) =

dzc =

Im [s1 ]
,
tan [] (s1 zc )]

and the compensator zero and pole are located at

zc = Re [s1 ] dzc ,

dpc =

Im [s1 ]
tan [] (s1 pc )]

pc = Re [s1 ] dpc

(21)

(22)

(23)

Examination of (21) shows that ] (s1 zc ) ] (s1 pc ) = ]Gc (s1 ) as it must. Since ]Gc (s1 ) is computed to provide
the proper phase shift at s1 to make the root locus pass through that point, the compensator parameters in (23) provide a valid
solution to the design problem. This solution is not necessarily any better than any other solution other than the fact that it
optimizes the ratio zc /pc .
Example 10:
From Example 5, the compensator must provide 36.24 at the point s1 = 4 + j5.4575. The phase shift of s1 itself is
]s1 = 126.24 . From (21), the angles of the compensator zero and pole are 81.24 and 45 , respectively. The horizontal
distances from s1 to the zero and pole are dzc = 0.8411 and dpc = 5.4575. This places the zero zc at s = 4.8411 and the
pole pc at s = 9.4575. The value of for this polezero combination is = zc /pc = 0.5119. Figure 7 shows the geometry
of this solution.

8) Multi-Stage Compensation: In the examples and discussion thus far, it has been assumed the compensator had one
pole and one zero. In some cases it may be desirable or necessary to use multiple stages of compensation. If the required
compensator phase shift ]Gc (s1 ) is very large, then the left-most allowed location for the compensator zero may be close
to the j axis or even in the right-half plane. Moving the zero to the left of this point means that one zero will not provide
enough positive phase shift, so that two or more zeros will be required.
Even with moderate values of ]Gc (s1 ) it may be desirable to use multiple stages of compensation. This would allow the
zero and pole to be moved farther to the left, away from the dominant pole location. In this way, the equations that apply to
second-order systems might be more applicable to the higher-order system.
The easiest way to design a multi-stage compensator is to assume that each stage provides the same amount of phase shift
at s1 . Since the phase shift of a product of complex numbers is the sum of the individual phase shifts, using this approach
means that the phase shift of each stage of the compensator is the total compensator phase shift divided by the number of
stages. The most usual case would probably be two stages; therefore, each stage would provide one-half of the total required
phase shift.
Once the total phase shift is divided by the number of stages, the design of each stage follows the procedure discussed in
the previous sections. The location of the compensator zero is selected to provide a phase shift larger than the phase shift
required by an individual stage. The compensator pole is computed to provide the correct phase shift for an individual stage.

14

Lag Compensated

Lead Compensated

10

3
5
s1

Imag Axis

Imag Axis

s1
0

1
2

3
4
8

2
0
Real Axis

10
10

1.4

1.2

2000

4000

6000

Fig. 6.

0.6

0.4

10000

0.2

4
6
Time (s)

0.8

8000

12000

Lead Compensated

2000

Step Response

Step Response

Lag Compensated

4
2
Real Axis

10

Comparison of lag and lead compensation for a particular system.

4
6
Time (s)

10

15

Alternative Placement of Compensator Pole and Zero


8

s1

Imag Axis

0
pc

zc

4
10

Fig. 7.

4
Real Axis

Selecting the compensator zero and pole to maximize = zc /pc .

Qnum_stages
The total compensator numerator then has the form i1
(s zci ), and similarly for the compensator denominator.
The gain Kc is then computed to satisfy
the
magnitude
criterion.
In
implementation, the gain can be assigned equally to each

stage by letting Kcstage = num_stages Kc .


Example 11:
The compensator phase shift computed in the previous examples is ]Gc (s1 ) = 36.24 . Although this compensator can be
designed with a single stage of compensation, it was shown in Example 8 that there was a restricted range of locations for
the compensator zero that allowed both the overshoot and settling time specifications to be satisfied. To use a single stage of
compensation, the zero must provide more than 36.24 at s1 .
In order to move the additional closed-loop poles far to the left of s1 , a two-stage compensator will be designed. This means
that each stage of compensation must provide 36.24 /2 = 18.12 at s1 . The phase angle for the zero of each stage will be
chosen as ] (s1 zc ) = 19 . Therefore, the angle of each pole must be ] (s1 pc ) = 0.88 . The horizontal distances from
s1 to the compensator zero and pole are 15.85 and 355.1, respectively. Therefore, the zero and pole are zc = 19.85 and
pc = 359.1. Using two poles and zeros at these locations, the gain is Kc = 2071.9, so the compensator is

2
2071.9 (s + 19.85)2
45.52 (s + 19.85)
Gc =
=
(24)
2
(s + 359.1)
(s + 359.1)
Using this compensator with the system G(s) = 8/ [s (s + 4)] produces an overshoot of 11.6%, which is less than three of
the single-stage compensators. The settling time is approximately the same as the first four designs presented in Example 8.
With the two-stage compensator, there are now four closed-loop poles, rather than three as before. Two of them are at s1 and

16

its complex conjugate. The remaining two closed-loop poles are at s = 357.1 j123. Although the overshoot is still a bit
higher than specified, it is smaller than the other designs that placed the compensator zero to the left of s1 , and the additional
closed-loop poles are much farther to the left of s1 than in any of the single-stage designs.

III. D ESIGN E XAMPLE


Both phase lead and phase lag examples will be presented in this section. The lead example will be presented first since that
is the compensator type most often used because it increases the system damping. After that, the lag compensator example
will be given. It will be seen that the design procedures for the two types of compensators are identical; just the role of the
pole and zero are reversed.
A. Phase Lead Example
1) Given System and Specifications: The open-loop transfer function for the system to be controlled is
Gp (s) =

0.375(s + 0.8)
s(s + 0.2)(s + 1)(s + 1.5)

(25)

The system is Type 1, so it will have zero steady-state error for a step input and a non-zero, finite steady-state error for a
ramp input. If unity feedback is placed around Gp (s), the closed-loop poles are located at s = {0.054 j0.4405, 0.9004,
1.6916}. Therefore, the original system is closed-loop stable with unity feedback and Gc (s) = 1. However, the damping
ratio of the complex conjugate pair of closed-loop poles is = 0.122, which corresponds to 68% overshoot for the standard
second-order system. The settling time is dominated by those complex conjugate poles since they are the closest to the j
axis, and the predicted settling time is 74 seconds. The uncompensated (plant with unity feedback) step response is shown in
Fig. 8, and it is clear that the predictions on overshoot and settling time are accurate for this system.
The specifications that are imposed on the system are:
overshoot in the response to a step input: P O 20%;
settling time for the response to a step input: Ts 16 seconds.
Thus, the compensator needs to reduce the overshoot and the settling time significantly. This requires that the effective damping
of the system be increased, which requires a phase lead compensator.
2) Selection of the Dominant Closed-Loop Pole: Using the equations for the standard second-order system, the damping
ratio that corresponds to an overshoot of 20% can be computed from (11), and is = 0.4559. The angle and slope of the radial
line associated with this value of are = 62.87 and 1.952, respectively. The settling time specification requires the dominant
closed-loop poles to have a real part computed from (13) of Re [s1 ] = 4/16 = 0.25. Combining these requirements places
the desired closed-loop pole at s = s1 = 0.25 + j0.488.
The uncompensated root locus is shown in Fig. 9, along with the point s1 . It is clear from the plot that the dominant branches
of the root locus need to be moved to the left in order to pass through s1 , again indicating the need for a lead compensator.
3) Designing the Compensator: The first step in designing the compensator is to determine the phase shift of Gp (s) at
s = s1 . This angle is ]Gp (s1 ) = tan1 [0.488/0.55] tan1 [0.488/ 0.25] tan1 [0.488/ 0.05] tan1 [0.488/0.75]
tan1 [0.488/1.25], so ]Gp (s1 ) = 41.58 (62.87 + 180 )(84.15 + 180 )33.05 21.33 = 225.77 = 134.23 .
Therefore, in order for s1 to be on the root locus, the compensator must provide a phase shift of ]Gc (s1 ) = 45.77 at s = s1 .
This can be done with a single polezero pair.
In order to use a single stage of compensation, the compensator zero must provide more than ]Gc (s1 ) = 45.77 at s1 , so the
left-most location for the zero is s = 0.725. Following the rule of thumb that the zero should be at or to the left of the second
real-axis open-loop pole, the compensator zero will be placed at the second pole, namely at s = 0.2. This location produces
an angle of 95.85 at s1 , so the compensator pole must have an angle at s1 of ] (s1 pc ) = 95.85 45.77 = 50.08 .
The distance from the real-axis projection of s1 to the compensator pole needed to provide this angle is equal to dpc =
Im [s1 ] / tan [50.08 ] = 0.4083. Therefore, the compensator pole is located at s = 0.6583. At this stage in the design, the
compensator is Gc (s) = Kc (s + 0.2) / (s + 0.6583).
The compensator gain is determined from the magnitude criterion |Gc (s1 ) Gp (s1 )| = 1. The gain calculation is
Kc =

|s1 | |s1 + 1| |s1 + 1.5| |s1 + 0.6583|


= 1.5192
0.375 |s1 + 0.8|

(26)

Note that the plant pole and compensator zero at s = 0.2 have been omitted from the calculation since they would cancel
out exactly. The final lead compensator for this example is
Gc (s) =

1.5192 (s + 0.2)
(s + 0.6583)

(27)

17

Uncompensated Step Response


1.8

1.6

1.4

Output Amplitude

1.2

0.8

0.6
PO = 67.6%
T = 72.6 s
s

0.4

0.2

Fig. 8.

10

20

30

40

50
Time (s)

60

70

80

90

100

Step response of the uncompensated system for the phase lead design example.

The compensated root locus and step response are shown in Fig. ??. The overshoot is 19.4%, and the settling time is 15.7
seconds, so the specifications have been satisfied. For the particular Gp (s) in this example, the dominant closed-loop pole
location s1 is sufficiently farther to the right than the other closed-loop poles to make the second-order equation for overshoot
hold for this higher-order system.
The following table summarizes the performances of the original system and the final compensated system. It is seen that
the compensator defined in (27) allows both of the transient performance specifications to be satisfied.
PO
Ts
Tr
essramp
closed-loop poles

Uncompensated
67.6%
72.6 sec
2.60 sec
1
0.054 j0.4405
0.9004
1.6916

Compensated
19.4%
15.7 sec
2.98 sec
2.17
0.25 j0.488
0.2
0.8285
1.8298

Specification
20%
16 sec
None
None
s1 = 0.25 + j0.488
None
None
None

Note that one of the compensated closed-loop poles is at s = 0.2, the location of one of the plants open-loop poles. The
reason for this is the fact that the compensator zero was placed at that same location. A compensator zero and a plant pole
(or vice versa) at the same location always results in a closed-loop pole at that location also. That pole does not affect the
settling time since there is a closed-loop zero at that point also. Two of the closed-loop pole for the compensated system are
at s = s1 and its complex conjugate as desired. The lead-compensated system has larger steady-state error for a ramp input
than the uncompensated system, but since that characteristic was not specified, this does not cause a problem. If there was a
specification on steady-state error, it would be considered at this point in the design process, using the technique described in

18

Uncompensated Root Locus


2

1.5

Imag Axis

0.5

0.5

1.5

2
3

Fig. 9.

2.5

1.5

1
Real Axis

0.5

0.5

Root locus of the uncompensated system and the desired closed-loop pole s1 = 0.25 + j0.488.

my notes Compensator Design for Steady-State Error Using Root Locus.


B. Phase Lag Example
1) Given System and Specifications: The system to be controlled has the open-loop transfer function
Gp (s) =

2(s + 1)
(s + 2)

(28)

The system is Type 0 so it will have a non-zero, finite steady-state error for a step input. Specifically, ess = 0.5 from Eqs.
(6) and (7). If unity feedback is placed around Gp (s), the closed-loop pole is at s = 1.33. Therefore, with Gc (s) = 1 the
closed-loop system is stable.
The specifications that are imposed on the system are:
overshoot in the response to a step input: P O < 5%;
settling time for the response to a step input: Ts 2 seconds;
steady-state error for a step input: ess = 0.
The steady-state error specification requires a Type 1 system (or higher). Therefore, the compensator will have to have a pole
at the origin as its minimum configuration. For purposes of designing the complete compensator, this pole may be included
with the plant model to form the following augmented model corresponding to Eq. (8).
G(s) =

2(s + 1)
s(s + 2)

(29)

19

Lead Compensated Step Response


1.4

1.2
Lead Compensated Root Locus
2
1

1.5

Output Amplitude

Imag Axis

1
s1

0.5
0
0.5

0.8

0.6
PO = 19.4%

1
T = 15.7 s
s

1.5
2
3

0.4
2

1
Real Axis

1
0.2

Fig. 10.

10
15
Time (s)

20

25

Lead compensated root locus and step response for the design example.

The transfer function in (29) is the one that will be evaluated relative to the transient performance specifications for the design
of the compensator.
2) Selection of the Dominant Closed-Loop Pole: Once again, the equations for second-order systems may be used to choose
the location of the dominant closed-loop pole. The overshoot specification imposes the constraint on the damping ratio of
> 0.6901, and the settling time specification imposes the constraint Re [s1 ] 2. For convenience, the damping ratio will
be = 0.707 since that corresponds to an angle of 45 and an overshoot of 4.3%. The dominant closed-loop pole will be
placed at s1 = 2+ j2. For the standard second-order system, this choice for s1 would allow both of the transient performance
specifications to be satisfied. Although the system in (29) is not the model of (9), and will be further modified by any additional
compensation other than a gain, we will use this choice as at least a starting point.
Since the root locus for G(s) lies entirely on the real axis (for both positive and negative K), the point s1 is not on the
root locus. Therefore, the compensator will need to provide phase shift at s1 in order to make a branch of the root locus pass
through that point.
3) Designing the Compensator: The first step will to compute the phase shift of the augmented system at s1 . This value is
]G (s1 ) = ] (s1 + 1) ]s1 ] (s1 + 2)
= 116.57 135 90
= 108.43 = 251.57
In order to place s = s1 on the root locus, the compensator must provide a phase shift at s1 of

(30)

20

]Gc (s1 ) = 180 (108.43 )


= 180 251.57
= 71.43

(31)

Since ]Gc (s1 ) < 0, phase lag compensation is needed, so the compensator pole will be to the right of the compensator
zero. The method described in Section II-E.7 will be used to choosing the pole and zero so that the value of = zc /pc will
be as small as possible. The phase angle of the point s = s1 is ]s1 = 135 , so the necessary angles of the compensator pole
and zero are
135 + (71.43 )
(32)
= 31.72
2

135 (71.43 )
] (s1 pc ) =
= 103.29
2
The distances from the real-axis projection of s1 to the zero and pole and the locations of the compensator zero and pole are
] (s1 zc ) =

dzc

dpc

2
= 3.2361
tan [31.72 ]
2
= 0.4721
tan [103.29 ]

(33)

and
zc = 5.2361,

pc = 1.5279

(34)

so the total compensator at this point in the design (including the pole needed for the steady-state error specification) is
Gc1 =

Kc1 (s + 5.2361)
s (s + 1.5279)

(35)

The gain Kc1 is computed from the magnitude criterion for the series combination of the original plant model in (28) and
the compensator in (35). The value of the gain is Kc1 = 0.6833, so the final compensator is
Gc1 =

0.6833 (s + 5.2361)
s (s + 1.5279)

(36)

A second compensator will be designed before the results of this first compensator are evaluated. Then a comparison of the
two designs will be made. Since ]Gc (s1 ) < 0, and the total phase shift of the compensator is the angle of the numerator
minus the angle of denominator, this compensator can be implemented without a zero, using only a pole (in addition to the
one at s = 0) and a gain. If this is done, the phase angle of the numerator is 0, so the phase angle of denominator is the total
phase shift of the compensator.
With this approach, ] (s1 pc2 ) = 0 ]Gc (s1 ) = 71.43 . The distance from the real-axis projection of s1 to the
compensator pole is dpc = 0.6667, so the pole is located at s = 2.6667. From the magnitude criterion, the gain is computed
to be Kc2 = 2.6667, so the final form of this compensator is
Gc2 =

2.6667
s (s + 2.6667)

(37)

Both compensators produce third-order systems when placed in series with Gp (s), and s = s1 is a closed-loop pole with
each of the designs. Figure 11 shows the compensated root locus plots for each of these designs. The small triangles on the
plots are the actual closed-loop poles. It is clear that the design parameter of having s1 be a closed-loop pole is satisfied. Two
points should be noted when comparing these root locus plots. The first point is the differences in the numbers of asymptotes
in the plots. The closed-loop system with Gc1 (s) has only 1 more pole than zero (relative degree = 1), so the only asymptote is
180 . The system with Gc2 (s) has 2 more poles than zeros, so the asymptotes have angles of 90 . This may be an important
difference if something causes the plant or compensator gain to increase in value. This would reduce the effective damping
ratio, leading to more overshoot. Although the closed-loop system would remain stable with this Gp (s) for all K > 0, this
would not be true for other systems. If the second compensator design increased the relative degree of a system from 2 to 3
(rather than from 1 to 2 as in this example), increases in gain would lead to an unstable closed-loop system if those increases
were sufficiently large. Therefore, the presence of the zero in Gc1 (s) provides additional stability robustness for the closed-loop
system.

21

Using Lag Compensator G (s)

Using Lag Compensator G (s)

c1

c2

3
s1

s1

Imag Axis

Imag Axis

1
2
3
4
12

Fig. 11.

10

6
4
Real Axis

5
3

1
0
Real Axis

Comparison of root locus plots with two lag compensator designs.

The second point to note is the location of the third-closed loop pole for the two designs. With Gc1 (s), that pole is very
close to the open-loop zero at s = 1. Therefore, its effect on the transient response will be small, even though the pole is
closer to the j axis than s1 . By contrast, the third closed-loop pole with Gc2 (s) is farther away from the open-loop zero and
also even closer to the j axis. The effect of this pole on the transient response might be significant.
Figure 12 shows the step responses for these two compensator designs, as well as for the closed-loop response of the
augmented system in (29). The only system that satisfies the settling time specification is the one with compensator Gc1 (s).
That system also has a much shorter rise time than the other two designs.
The following table summarizes the results from this example.
PO
Ts
Tr
essstep
closed-loop poles

Gp (s)
16.67%
2.12 sec
undefined
0.5
1.33

G(s)
0%
5.50 sec
2.67 sec
0
0.5858
3.4142

Gc1 (s)Gp (s)


0.65%
1.075 sec
0.749 sec
0
2 j2
0.8944

Gc2 (s)Gp (s)


0%
4.66 sec
1.88 sec
0
2 j2
0.6667

Specification
< 5%
2 sec
None
0
s1 = 2 + j2
None

The overshoot and settling time for the uncompensated system are based on the final value for the output of 0.5. Since the
uncompensated system has the same number of poles and zeros, there is an impulse in the step response, and the maximum
value (0.6667) occurs at t = 0. This value was used in the computation of overshoot. Since the output never takes on a value
equal to 10% of its final value, the rise time is undefined.
Augmenting the system with the pole at the origin satisfies the steady-state error and overshoot specifications, but not the
settling time specification. As previously mentioned, the system using Gc1 (s) satisfies all of the specifications. The second
compensator design fails on the settling time specification due to the closed-loop pole at s = 0.6667.
For the augmented system of (29) and the transient response specifications in this example, gain compensation can be used.
Both the specifications are satisfied for all Kc 4.55. With this approach, the total compensator would be Gc (s) = Kc /s. The
point s1 would not be a closed-loop pole; both poles would be real. Very short rise and settling times can be achieved by using
large gain values. However, a potential drawback with this approach is large acceleration in the system output. The control
signal would change values rapidly near the beginning of the response causing the output to change rapidly also, with the
output signal approaching a discontinuity for large gains. If the output signal represented mechanical position of some object,
a large acceleration might be objectionable. The rapid change in the value of the control signal might also exceed physical
limitations of the actuator providing motion to the system. The control signals generated by Gc1 (s) and Gc2 (s) change more
smoothly and limit the acceleration.
This example illustrates the design technique of satisfying transient performance specifications using root locus. Even though
the compensated system was third-order and contained one zero, the selection of the point s1 yielded a compensator design
Gc1 (s) that more than satisfied the specifications. If there was a specification on steady-state error for a ramp input, it would

22

Lag Compensated Step Responses


1.4

1.2

Output Amplitude

0.8

0.6

0.4

0.2

G (s); T = 1.08 s
c1

Gc2(s); Ts = 4.71 s
G (s) = 1/s; T = 5.5 s
0

Fig. 12.

5
Time (s)

10

Comparison of closed-loop step responses for two lag compensators.

be considered at this point in the design process, using the special lag design described in my notes Compensator Design
for Steady-State Error Using Root Locus.
R EFERENCES
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]

J. DAzzo and C. Houpis, Linear Control System Analysis and Design. New York: McGraw-Hill, 4th ed., 1995.
R. C. Dorf and R. H. Bishop, Modern Control Systems. Reading, MA: Addison-Wesley, 7th ed., 1995.
K. Ogata, Modern Control Engineering. Upper Saddle River, NJ: Prentice Hall, 4th ed., 2002.
G. Franklin, J. Powell, and A. Emami-Naeini, Feedback Control of Dynamic Systems. Reading, MA: Addison-Wesley, 3rd ed., 1994.
G. Thaler, Automatic Control Systems. St. Paul, MN: West, 1989.
W. A. Wolovich, Automatic Control Systems. Fort Worth, TX: Holt, Rinehart, and Winston, 3rd ed., 1994.
J. V. de Vegte, Feedback Control Systems. Englewood Cliffs, NJ: Prentice Hall, 3rd ed., 1994.
B. C. Kuo, Automatic Controls Systems. Englewood Cliffs, NJ: Prentice Hall, 7th ed., 1995.
N. S. Nise, Control Systems Engineering. New York: John Wiley & Sons, 3rd ed., 2000.
C. Phillips and R. Harbor, Feedback Control Systems. Upper Saddle River, NJ: Prentice Hall, 4th ed., 2000.
G. C. Goodwin, S. F. Graebe, and M. E. Salgado, Control System Design. Upper Saddle River, NJ: Prentice Hall, 2001.

S-ar putea să vă placă și