You are on page 1of 12

IN 227 Control Systems Design

Lecture 19

Instructor: G R Jayanth
Department of Instrumentation and Applied Physics
Ph: 22933197
E-mail: jayanth@isu.iisc.ernet.in

Control of non-minimum phase systems

In all the control design techniques we have discussed, the plant that we intend to control took a back
seat. Beyond the fact that it imposed some limits on our ambitions as control engineers, we assumed that
it caused no further inconveniencesthere was a significant bandwidth, viz., the control bandwidth,
which could be made arbitrarily large ,wherein we could maximize the loop gain by any desired amount
and thus achieve the goal of control, namely x(t)=r(t).
The plant was intentionally assumed to be cooperative because we wanted to shine the spotlight on the
different control techniques that could be employed to control a general plant. Cooperative plants possess
all their poles and zeros in the left half of the complex plane and do not possess any delays. However,
there exist a class of linear plants that refuse to be cooperative and demand special attention by virtue of
the inconveniences they cause to conventional controllers: they impose fundamental theoretical limits to
the achievable loop gain and the gain cross-over frequency. This lecture is devoted to studying the control
of such plants.
These plants could be afflicted with three types of maladies (a) time delay (b) right-half plane zeros and (c)
unstable dynamics.
There is not much connecting these three types of plants except that they are all non-minimum phase
plants. For minimum-phase plants, when we specify the gain of a plant, we also uniquely fix its phase.
However, a non minimum-phase plant possesses lot higher phase lag for the same gain characteristics.
Example: Consider a minimum phase system 1/(j+1) and a non-minimum phase system 1/(j-1). The
magnitude bode plots for both are exactly the same: they starts at 0 dB and begin to roll off at
20dB/decade beyond =1rad/s. However, when we look at the phase plots, the former starts at 0 deg and
tends to -90 deg. at higher frequencies, while the latter begins at -180 deg. and ends at -90 deg.
This lecture is devoted to studying the control of plants with delays and with right half plane zeros.
When we are dealing with such weird plants, all our traditional compasses of stability in Bode plot,
namely the gain and phase margins, either break down or may require to be interpreted with care. Hence
in analyzing the control of non-minimum phase systems, we step back and check for stability using Nyquist
stability theory, before we transfer the system to the familiar domain for design proposes, viz., Bode plots.

Control of systems with time-delay

Time-delay is ubiquitous in process industries and other systems where it takes finite
time for a particular event to reach the sensor. This may be because the event-to-besensed may take time to physically or chemically move to the sensor. For example, in
a steel plant, the thickness of a steel sheet can be measured only after it has cooled
down, but the thickness can be controlled only when it is hot. Thus, there is a finite
time lag between control and measurement during which the sheet cools. Of course,
one may choose to be more philosophical and say that time delays are inescapable
facts of life thanks to Einsteins special theory of relativity. Whether the delay is of r
practical consequence or not is a matter that engineers have to decide.
Assume that a certain system has a time delay T in addition to certain dynamics P0(s).
Since the Laplace transform of a delay is e-sT, the transfer function of the plant
without delay (P0(s)) gets multiplied by the term e-sT. Thus, the control block-diagram
of a 1 DOF control system with delay appears as shown on the right.
It is not intuitively clear from the block-diagram why the time delay could cause any
inconvenience. Thus, to understand this, we turn to the mathematical tools, in
particular, Nyquist stability theory.
As explained before, we first check if Nyquist stability criterion can be uncritically
applied: the closed-loop transfer function is X/R=CP0e-sT /(1+CP0e-sT). The
denominator is 1+CG/esT. The zeros of esT which are the poles of the open-loop
system exist at Re(s)-. Thus, if P0 (s) is stable, then there are no open-loop poles
CP0e sT 0 . Thus, we can employ the same Nyquist path
on the RHP. Furthermore, lim
s
(C) as we did for ordinary P0 (s) , viz., a semicircle centered at the origin
encompassing the entire RHP. All the points on the curved part of C collapse to the
origin. For stable closed-loop system, we need to evaluate CP0e-sT along s=j.

d
u

C ( s)

P0 (s)e sT

C
R

Control of systems with time-delay

Since we know how to plot CP0(j), we focus on the effect of e-iT. We note that e-iT
=cosT-isinT. Thus, the magnitude of e-iT is one, regardless of frequency. The phase is CP (j )e-i T
0
1
1
tan-1(-sinT/cosT)= -T. Since e-iT multiplies CP0(j), we see that at each frequency ,
the effect of e-iT is to rotate the phasor CP0(j) by an angle T without affecting the
=0
=
gain. At frequencies close to zero, the effect is barely noticeable. At higher frequencies,

0
-1
the phasors get rotated by larger angles. For >>2/T, the phasor gets rotated several
CP0 (j)
-T
cycles. In particular, we note that the phase of the transfer function CP0 e-iT becomes 180 at a lesser frequency 1 than that of CP0 (which is defined to happen at 0). The gain
1

of CP0 at 1 however, is much higher than the gain at 0. Thus, a time delay has the effect
of reducing the gain margin of the overall system. If the delay T is larger, then the phasor
CP0 (j)e-iT
at an even lesser frequency gets rotated to align along the negative real-axis. At such a
Bode plot of P0 (s)=100/(s+10)
small frequency, the gain would in general be even greater. If T is sufficiently large, then it
will rotate a phasor at such a small frequency that at that frequency the plant gain is
greater than 1. In this situation the system becomes unstable. Thus, we see that delays can
destabilize closed-loop systems that were originally stable.
Since we can apply Nyquist stability principles exactly as before for a system with time
delay, we can safely transfer it to the Bode domain. Here the effect of the delay is to add a
phase lag T without affecting the gain curve (since |e-iT|=1). Since Bode plots are log
plots, we see that the phase of e-iT is an negative exponential function.
For the sake of illustrating the influence of this phase lag, we choose a first order system
P0(s)=100/(s+10). We see that the phase of this function tends to -90 as . Thus, this
Bode plot of G(s)=100e-0.01s/(s+10)
20
plant possesses no phase cross-over and thus can be designed to achieve high gain over
10
arbitrarily large bandwidths.
0
However if we add a delay of just 0.01s, i.e., P0(s)=100e-0.01s/(s+10), the phase of the
-10
open-loop system crosses -180. We now have to ensure that the gain of the open loop
10
10
10
10
0
system is less than 0dB at this frequency . Thus a system which can be designed to possess
-100
arbitrarily high bandwidth without delay gets limited in bandwidth due to the delay.
-180
-200
Bode plots make it abundantly clear how just phase lag addition can limit control
-300
performance.
10
10
10
10
Bode Diagram

10
0

-10

-20
0

Phase (deg)

Magnitude (dB)

20

-45

-90
-1
10

10

10

10

10

Frequency (rad/sec)

-1

-1

Control of systems with time-delay

If the delay is small enough that the phase cross-over frequency is close to the gain cross-over frequency, we
can continue to use the same design tools as for conventional control systems in order to control the plants.
However, since we traditionally deal with rational functions of , and e-iT is not a rational function, it would be
convenient to approximate e-iT by rational functions for design purposes. However, unlike e-i whose phase
tends to negative infinity, the phase of any rational function will settle at some integral multiple of -90. Thus,
we need to note that rational functions can approximate e-iT only for finite frequency range.
We note that truncated Taylors series expansion of e-iT (as 1- T+(T)2/2!-..up to nth degree) will not work
because any truncation will result in polynomial of whose magnitude also varies along with its phase This is
a poor approximation of e-iT because the magnitude of e-iT is constant.
Pade approximants: The most popular approximants for e-iT were proposed by H. Pade. The general nth
degree approximant is given by
sT ( sT ) 2 n(n 1)
( sT )3 n(n 1)(n 2)
1

...
2
2! 2n(2n 1)
3! 2n(2n 1)(2n 2)
P( s )
sT ( sT ) 2 n(n 1)
( sT )3 n(n 1)(n 2)
1

...
2
2! 2n(2n 1)
3! 2n(2n 1)(2n 2)

sT ( sT ) 2
sT
1

2
12 and so on.
2
For n=1, we see that P( s)
. For n=2, we have P( s)
sT
sT ( sT ) 2
1
1

2
2
12
1

We see that Pade approximants, unlike, Taylors series truncation, yield rational functions with unity gain at all
frequencies, and thus, correctly reflect the effect of e-iT on the magnitude plots. Depending on the extent of
n
sT
accuracy required, we use one of the approximants in plotting phase plots.
1
It is worth pointing out that other approximants are possible. For example, Q( s) 2n n is another valid nth
sT
degree rational function approximant for e-iT.
1
2n

Control of systems with time-delay

As we noted before, the use of conventional control techniques is fine if the phase
cross-over is not seriously affected by the delay. However, if it does get reduced
dramatically, conventional design techniques lead to very poor performance. This is
illustrated below.
Example 1: Suppose in the previous delay-free plant that we chose (P0(s)=100/(s+10))
we incorporate a time delay of 2s, i.e., P0(s)=100e-2s/(s+10). We are required to
develop feedback controllers for this plant.
Solution: The bode plot of the open-loop system shows that the phase cross-over
frequency is significantly less than the gain cross-over frequency. We also note that
the gain margin<0. Thus, we need to reduce the gain to less than 1 at the phase crossover frequency.
We can choose to do this using a proportional controller but, in doing so, we will also
sacrifice low-frequency performance. Thus, we use integral control. This reduces the
phase further by 90. Since the slope of the phase curve at the phase cross-over is
large, the addition of -90 does not significantly reduce the phase cross-over frequency.
Thus, we do not bother to cut off the integrator characteristic with PI-control.
An integral controller with gain 0.041, i.e., C(s)=0.041/s, will ensure that the gain
margin >0. For this controller we see that the phase margin is >0. However, in using
this control, we have dramatically reduced the control bandwidth (by more than a
factor of 100!). Thus, the step response shows very slow transience (transience slower
than the speed of the open-loop system). Furthermore, since the closed-loop system
is given by X/R={CP0/(1+CP0e-sT)}e-sT, we note that we will always have a delay T in the
response of the system.
We see that feedback controllers perform pretty poorly in the presence of delays.
Further, there appears to be no way for us to substantially improve the performance.
We shall soon see that the performance of these systems is theoretically limited.

Gain cross-over
20
10
0
-10
-1
10

10

10

10

0
-100

-180

-200
-300
-1
10

10

10

10

Phase cross-over
50
0
-50
-100

10

10

0
-100
-200

PM

-300

10

10

Log()
1.4
1.2
1

x(t) 0.8
0.6
0.4
0.2
0

10

20

Time (s)

30

Control of systems with right half plane zeros

20Log(Mag.)

A nonminimum phase plant, i.e., one with a right half plane (RHP) zero has the following form: PNMP(s)=(a-s)P1(s),
where, P1(s) does not contain any non-mimimum phase terms.
The only way to study the effect of a transfer function on closed-loop stability is by means of Nyquist plots. We note
the plant PNMP(s) has no poles in the RHP. The presence of a zero of PNMP(s) in the RHP does not in any way cause
problems for applying Nyquist stability theory to search for closed-loop poles in the RHP. Thus, the rules remain
unchanged: the system is stable if CPNMP(j) does not encircle -1. Therefore, Bode plots can also be used without any
modification.
The way a non-minimum phase term would affect the closed-loop dynamic can be seen by writing PNMP(s)=[(as)/(a+s)][PMP(s)] where, PMP(s)=P1(s)(a+s). We know how to draw the bode plot of the minimum phase part, viz., PMP(s).
We therefore examine the other term (a-s)/(a+s). This term, and the product of such terms (if there happen to be
multiple RHP zeros) are called Blaschke products.
We note that the gain of this term is 1 at all frequencies. Thus the Blaschke product does not affect the magnitude
characteristics of the plant. However, the phase of this term is given by -2tan-1(/a): it starts at 0 at =0, becomes -90
at =a and tends to -180 as . Thus, the term adds a phase lag of -180, with the corner frequency being =a .
Since all causal physical systems display increased phase lag with frequency, a RHP zero mandatorily causes the phase
of the system to cross -180 a bit beyond =a . Thus, an RHP zero affects control performance the same way that timedelay does-it reduces the phase cross-over frequency.
Note that we cannot cancel an RHP zero with a controller pole at =a. This is because in practice, the cancellation can
never be perfect. Thus, a branch of the root locus would exist between the zero and the pole, and as the result, the
closed-loop system will be unstable.
a
Log()
Slope=-20dB/decade
Bode plot of (a-s)/(a+s)

hase (deg.)

~0.1a
0
-90
-180

~10a

Fundamental upper limit to the gain cross-over frequency for


systems with RHP zeros
20log |L|

20log |L

| (=20log |L |)

NMP
MP
In order illustrate fundamental limits to achievable benefits of feedback imposed by the
nonminimum phase plant, we shall consider the design of a control system where we have
-40 dB/decade
specified that the loop gain should possess a phase margin of PM radians and a certain gain
log
gc
margin GM dB.
GM
It is assumed first that a minimum phase, i.e., well behaved controller C(s) is used to
control the nonminimum phase plant PNMP(s). The resulting loop gain is given by LNMP=CPNMP.
log
We need to shape LNMP to achieve the desired performance and stability specifications. For
reference, we shall also consider the minimum phase counterpart LMP=CPMP whose only L
NMP
difference from LNMP in that it is not multiplied by the Blaschke product (a-s)/(a+s). Since the
PM
-
pc
Blaschke product does not affect the magnitude characteristics, LNMP has the same
magnitude characteristics as that of the minimum phase counterpart. This is sketched on the
right. The phase of the loop gain LNMP at any frequency , however, is lesser than the phase
of LMP by the amount 2tan-1(/a), i.e., LNMP LMP 2 tan 1 ( a)
Let LNMP cross 0 dB at gc and let the slope of the magnitude characteristic in the
neighborhood of gc be -40 db/decade. Bodes gain-phase relationship tells us that the
LNMP LMP 2 tan( a)
phase of the minimum phase loop gain LMP at any frequency is determined predominantly
by the slope of the magnitude characteristic near alone, regardless of the actual shape of
the magnitude characteristic of LMP at other frequencies. Thus, the phase of LMP at gc is -
rad. Consequently, the net phase of the loop gain LNMP at gc isLNMP (gc ) 2 tan 1 (gc a)
To satisfy the prescribed phase margin PM, we note that LNMP (c ) PM . Equating these

two we get tan 1 ( gc ) (1 ) PM or, equivalently, gc cot PM


a
2
a
2

Since cot() is a decreasing function of , we note from the equation above that, for a given
PM, the gain cross-over frequency is a decreasing function of . Thus, the highest gain crossover frequency is obtained for the smallest value of . However, since the low frequency
loop gain needs to be greater than 0dB, we require that >0 in order for gain to cross the
0dB line.

Fundamental upper limit to the gain cross-over frequency for


systems with RHP zeros

Thus, for the given phase margin PM, the absolute maximum gain cross-over frequency is obtained when ~0 and is
given by gc a cot PM 2
max
Thus, we see that the maximum possible value of the gain cross-over frequency gc is theoretically limited to
a.cot(PM/2) due to the RHP zero at s=a.
Likewise, we can derive the theoretical limit to the maximum gain cross-over frequency for which the specified phase
margin PM and the gain margin GM are simultaneously satisfied. This is done below:
GM is the plant gain at the phase cross-over frequency pc , i.e., where the phase is -. Assuming that the slope of
LNMP remains approximately -40 db/decade between gc and pc, and thus, the phase of LMP remains at , we
have, 2 tan 1 ( pc a) or equivalently, pc a cot( 2)
Since the slope of LNMP remains approximately -40 db/decade between gc and pc, (fig. below right), we have
gc . Using the fact that
GM
pc a cot( 2) and gc a cot ( PM ) 2, we obtain, GM log tan ( PM ) 2
log

40

pc

40

tan( 2)

The last equation above is a transcendental equation in . It is solved numerically to obtain and subsequently used
to evaluate gc. The figure below (left) plots the values of gc (normalized with respect to a) for different phase and
gain margin requirements.
20log |LNMP|

-40 dB/decade
log

gc

gc

GM

a
LNMP

-
From Ref. [1]

log

pc

Fundamental limits to control performance

20log |L|
The analysis of the previous page indicates that for a specified stability margin GM
20log|LNMP()|
and PM, the maximum possible gain cross-over frequency is limited, typically to a
small fraction of a, where a is the location of the RHP zero. We shall now examine
the consequences of this limit on the maximum achievable loop gain.
log
We start by noting that if we want the phase margin to be PM, then the maximum
gc

phase-lag that we can allow for the loop gain LNMP is -+PM for all frequencies
<gc.
The phase lag of the Blaschke product at any frequency is -2tan-1(/a). Thus, the
maximum permissible phase lag for LMP therefore is -+PM +2tan-1(/a). We shall
log
-1
express this as a fraction max() of , i.e., -max()=-+PM +2tan (/a)
LNMP
Given the maximum permissible phase lag for LMP , Bode gain-phase relationship
PM
reveals the maximum negative slope that the gain characteristic of LMP can have at
-
the frequency is -40max()dB/decade, which is the same as that for LNMP .
Maximum permissible
For any frequency , the loop gain |LNMP()| will be maximum when the gain
phase lag
characteristic has the maximum permissible slope for all frequencies between

and gc. Thus, the maximum permissible value of |LNMP()| is given by 20log LNMP ( j) max 40 max (u)du

where u=log .
At frequencies <<a, 2tan-1(/a)~0, so that max()= -1+PM/. Thus, at low
frequencies, the slope of the gain characteristic reaches its maximum permissible
Low frequency slope =
constant value of -40(1-PM/). The schematic of the gain characteristic is shown on
-40(1-PM/) dB/decade
the right.
The consequences of the limit in maximum achievable loop gain at any given
20log |L|
Maximum
frequency are profound: for example it is impossible to reject disturbances near the
achievable loop
frequency 2 shown on the right due to the low loop gain there. Even within the
gain
control bandwidth, there is an upper limit to the achievable loop gain at any
frequency 1, and thus an upper limit to the extent to which we can reject
log
disturbances and achieve robustness against plant parameter variations.
gc 2
1
If a RHP zero is slow, i.e., is located close to the origin, the control bandwidth gets
drastically reduced regardless of how fast the plants poles might be.
gc

Control of systems with time-delay

We note that the non-minimum phase terms result in poor closed-loop


performance with respect to rejection of disturbances and parameter variations.
Good tracking is one place that open-loop control can help substantially. If we
simply invert the undelayed part of the plant by means of a feedforward controller r(t) f
f
e sT
(i.e., use the strategy schematically shown on the right) without using feedback
x(t)
Plant
control we can at least achieve excellent dynamic tracking performance, even if the
delay of T seconds between the input and the output continues to exist.
Now, if we know that when f-1() cannot be directly implemented, we can use high
+
u
e
gain-based open-loop model-inversion. The corresponding block-diagram, which r
e sT
f
h
x
was discussed in Lecture #5, is shown on the right.
However, we know that open-loop control is sensitive to drifts in plant parameters.
f
Thus, we use feedback only to correct for the plant inversion that was already
z
achieved by open-loop control. In other words, we compare the estimate of the
output of our feed forward controller (z) with the actual output (x) and the error is
Feed forward based input tracking
fed back. Thus, feedback plays a weak role in the overall control system: it is active
only when there is difference between the nominal models of the plant and the
actual one. The resulting structure of the control system is called Smith Predictor
or Smith Controller. We note that since conventional controllers did a poor job of
controlling a plant with delay, we are forced to model the plant as accurately as we
possibly can, and subsequently employ open-loop control to invert the model. Only
errors in modeling are compensated by means of feedback.
C ( s) P0 ( s)e sT
X (s)
The transfer function of the control system is given by R( s)
sT
sT
1

1 C ( s) P0 ( s) C ( s)[ P0e

P0e

1.5

r +
1

Comparison of
step responses

With smith predictor


0

10

20

30

40

Smith controller

P0 ( s)

50

e sT
z

P0 ( s)

Conventional control
0.5

C ( s)

sT

References
(1) Marcel Sidi, Design of Robust Control systems From Classical to Modern Practical Approaches, Kreiger
Publishing Co. FL, USA, (2001)
(2) G C Goodwin, S E Graebe, M E Salgado, Control System design, Prentice-Hall of India (2001)

You might also like