Control System Design Astrom
Control System Design Astrom
Introduction1
This first lecture has two parts. The first part gives an introduction and overview
of the course, starting from two examples of modern control, a DVD-reader and car
dynamics. The second part of the lecture is a brief review of linear input-output
models in continuous time used in the basic course. The concepts of signal norm
and system gain are introduced.
Track
Pit
0.74 m
Figure 1.1
The right picture shows pits forming tracks on the DVD surface.
The disk surface is reflective, so that laser light is reflected back. Data bits
are represented by pits of different lengths in tracks on the disk. These pits make
the laser beam interfere destructively with itself, and therefore the pits look black
to the laser.
The surface velocity is constant (about 3.5 m/s), meaning that the disc should
rotate at different speeds depending on the current reading position. The challenge
of the control problem is related to the fact that only 0.022 m deviations from the
1 Written
Lecture 1.
Introduction1
bit-track can be accepted. At the same time, a disk is always slightly asymmetric,
causing it to oscillate up to 100 m per rotation, and the rotation speed is up to 23
Hz (for single speed). The tracking controller must compensate for this oscillation.
A typical DVD player has a pick-up-head consisting of a laser, an astigmatic
lens, and a light detector with four fields see Figure 1.2. The lens is mounted
on springs in the axial (focus) and radial direction, and can be moved by electromagnets. This way, the laser spot can be moved very fast in a small range (a few
hundred tracks sideways). The lens and laser are mounted on the sledge, which
can move over the whole disk (in radial direction), but with much less precision
and speed.
Tracks
Radial electromagnet
Lens
Disk
A B
C D
Springs
Light detectors
Focus electromagnet
Laser
Sledge
Pickup head
Figure 1.2 The pick-up-head has two electromagnets for fast positioning of the lens (left).
Larger radial movements are taken care of by the sledge (right).
Four light detectors are available to estimate the focus error and radial error
of the lens. Measurements are taken with a sampling frequency of 40 kHz and
the DVD standard specifies that the speed of control (cross-over frequency) must
be at least 2.4 kHz.
It turns out that most of the main topics of this course are relevant for the
solution of the DVD control problem and we have therefore chosen to use it as a
demonstrator. The focus control will be treated in a lecture, while a lab exercise
towards the end of the course is devoted to disc track following.
Figure 1.3
1.2
V
r
V
r
0
b1
(u1 + u2 u3 u4 ) +
b2
b3
where V is lateral speed and r is angular velocity. There are five control signals,
the steering angle and the brake forces u1 , u2 , u3 and u4 on the four wheels.
U
r
Figure 1.4 A modern car relies on feedback control for comfort, safety and fuel efficiency.
The left picture shows a test-car used in a research project together with DaimlerChrysler.
The state is generally not available for direct measurement. Even if the angular
velocity of each wheel can be measured, there is always some discrepancy between
the rotational speed and the speed over ground. Hence the velocity of the car
must be estimated based on information from several sources and the remaining
uncertainty must be taken into account in the control algorithms.
A typical sampling frequency for speed measurements is a few milliseconds.
This may sound fast enough compared to typical car dynamics, but when the
purpose is to prevent wheel-lock or accidents, a delay of a few milliseconds can in
fact be a severe obstacle for proper control performance.
lateral velocity
brake forces
Vehicle
steering angle
Figure 1.5
yaw rate
Introduction1
Lecture 1.
u (t )
u (t )
t
Process
Hold
Sampler
yk
uk
uk
y (t )
D-A
Computer
A-D
yk
Idea/Purpose
Analysis
Experiment
Synthesis
Implementation
Figure 1.6
Everything starts with an idea about the purpose of the control task. In simple
cases, it is possible to directly come up with a solution proposal that can be tested
experimentally and be accepted, possibly after minor modifications. However, in a
vast number of applications costs and time can be reduced by analyzing or simulating a mathematical model before trying real experiments. The purpose of the
diagram is to illustrate this methodology. Note that the arrows point in two directions. Failure in the experimental phase could not only require reimplementation,
but also new analysis, more accurate models, or even redefinition of the control
purpose.
Imagine stepping through the diagram in order to design a controller for car
dynamics as in the previous example. Suppose that a controller has been synthe-
1.4
sized based on the given two state model. Implementing a controller on a prototype
car is costly, so a second step would typically involve computer simulation. For this
purpose a more complex and accurate car model is needed, a model that is less
transparent from a synthesis perspective but better suited to reveal the deficiencies of a proposed controller. If the simulations fail, a reason could be that the
two state model was too simple and that additional features need to be taken
into account in the synthesis phase. After a sequence of attempts, one could hope
to find a solution ready for experimental tests. Alternatively, persisting failures
could be an indication that the original goal was overly optimistic and impossible
to achieve.
The main emphasis of this course is on the analysis/synthesis phase of the
diagram in Figure 1.6. However, to keep the big picture in mind, there will also
be lectures and laboratory sessions devoted to modelling, implementation and
experiments.
The main topics of the course are the following
Lecture 1.
Introduction1
Note that the formula remains valid for multivariable systems, i.e. when both u(t)
and y(t) are vector valued.
The map from u to y is linear provided that x(0) = 0. Introducing the impulse
response (t) as
(t) =
(t )u( )d = [ u](t)
0.5
0.4
0.3
0.2
0.1
0
0.1
1
10
10
u j (t) = (t)
1
0.8
0.6
0.4
0.2
0
1
u(t) =
t<0
u0
t0
(t s)u0 ds =
Z
t
0
( )d
u0
1.4
yi (t)
0.8
0.6
0.4
0.2
0
1
10
10
u j (t)
1
0.8
0.6
0.4
0.2
0
1
The main use of the Laplace transform is however to characterize the frequency
response. The input u(t) = u0 sin t gives
y(t) =
( )u(t )d = Im
Z
t
0
( ) ei d ei t u0
yi (t)
0.2
0.1
0
0.1
0.2
10
12
14
16
18
20
10
12
14
16
18
20
1.5
u j (t)
1
0.5
0
0.5
1
1.5
Lecture 1.
Introduction1
Magnitude
10
10
10
Phase
45
90
135
180
10
10
Frequency (rad/sec)
It should be noted that each additional factor in the transfer function contributes
additively to the Bode plots:
log p G1 G2 G3 p = log p G1 p + log p G2 p + log p G3 p
arg G1 G2 G3 = arg G1 + arg G2 + arg G3
The Nyquist diagram is obtained by plotting G (i ) directly in the complex
plane for different values of :
0.6
Im
0.4
0.2
Re
arg G (i )
0
0.2
p G (i )p
0.4
0.6
1
0.5
0.5
30
0.25 dB
0.5 dB
20
1 dB
Amplitude
1 dB
3 dB
10
3 dB
6 dB
6 dB
10
12 dB
20
20 dB
30
40
360
40 dB
315
270
225
180
Phase
135
90
45
The level curves of p G /(1 + G )p and arg G /(1 + G ) are plotted as dotted lines to
support use of the diagram in controller design.
1.5
q yq2 :=
sZ
p y(t)p2 dt
According to a theorem known as Parsevals formula, the same norm can be defined in frequency domain as
q yq2 =
1
2
pLy(i )p2 d
For a system S with input u, output S (u) and zero initial state, the L2 -gain
is defined as the largest possible fraction between the input norm and the output
norm
qS q := sup
u
qS (u)q
quq
The system is called input-output stable (or L2 -stable) if its L2 -gain is finite. For
example, a time delay does not change the signal norm, so it has gain one. However,
an integrator has infinite gain, since an input u(t) that is identically zero for t 1,
can give an output y(t) that is a nonzero constant for t 1. Hence, the fraction
q yq2 /quq2 can be arbitrarily large.
More generally, the L2 -gain of a system can be obtained as the maximum
amplitude in the Bode diagram:
THEOREM 1.1
A stable system with transfer function G (s) has the L2 -gain
q G q := sup p G (i )p
Remark. For multivariable systems the p G (i )p should be interpreted as the matrix norm (the largest singular value) of G (i ). This case will be studied more
carefully later.
Proof. Let y be the output corresponding to the input u. Then
q yq2 =
1
2
pLy(i )p2 d
1
2
The inequality is arbitrarily tight when u(t) is a sinusoid near the maximizing
frequency.
2
Example 1
a. For a time delay G (s) = esT we have p G (i )p " 1.
b. For an integrator p G (i )p = p i1 p = 1 which is unbounded = 0.
c. The Bode diagram plotted in the previous section has a peak magnitude
about 0.5 at the frequency 2 rad/sec. Hence, the L2 -gain of the corresponding
system is smaller than one and the highest gain is obtained for an input sinusoid
of this frequency.
2
Lecture 2
Figure 2.1 Lawrence Sperry demonstrates a stabilizing gyroscopic controller. He waves his
hand in the air, while his mechanic is walking on the wing.
Another striking example was the construction of the first electronic feedback
amplifiers, that were necessary to build long distance telephone connections in
the 1930s. In this case, high gain feedback was needed to reduce the nonlinear
signal distorsion. Stability problems became a major issue, and the development
of frequency domain stability criteria was critical for successful implementation.
A modern example is the maneuver test for Mercesdes A-class, that created
unstable oscillations severe enough to turn the car over. The problem was solved
by introducing electronic feedback control.
Recall from the previous lecture that a system is called input-output stable
(or L2 -stable) if its L2 -gain is bounded. A transfer function is called stable if it
corresponds to an input-output stable system. The following stability criterion is
available for linear time-invariant systems.
THEOREM 2.1
A rational transfer function G (s) is stable if and only if all poles of G have negative
real part. In particular, if G (s) = C (sI A)1 B + D, it is sufficient that all
eigenvalues of A have negative real part.
1 Written
10
2.1
Figure 2.2
1 2
1
x =
+
u
3 2
0
y = [1 1] x + u
( I)
( I I)
y(t) =
e t u( ) y( ) d
( I ) The first system is input-output stable due to stable eigenvalues of the system matrix. A two-by-two matrix like this is stable if and only if the trace is
negative (here 3) and the determinant is positive (here 8). This is because
the trace is the sum of the eigenvalues and the determinant is the product.
In general, eigenvalues can be computed by the matlab command eig(A):
>> eig([-1 2; -3 -2])
ans =
-1.5000 + 2.3979i
-1.5000 - 2.3979i
Note that the coefficients of the characteristic polynomial should not be computed, at least for high order systems, since this generally leads to numerical
difficulties.
Y ( s) =
The transfer function 1/(s + 2) shows that the system is stable, since the
only pole 2 is negative.
2
Our main objective is to study stability of feedback loops. From the basic course,
we recall the Nyquist criterion, which supports understanding by graphical illustrations.
11
Lecture 2.
1
Am
0.5
G ( s)
Im
- n 6
1
0.5
c
1
1
0.5
0
Re
0.5
Figure 2.3 The closed loop system remains stable as long as the Nyquist diagram does
not encircle 1. The amplitude margin Am and phase margin m measure the distance from
instability.
u(t) = k[ x(t T ) + x (t T )]
The feedback loop is illustrated in the figure below. Nominal values of the parameters are k = 1, c = 1 and T = 0. Let us investigate how much margin there is in
each of the parameters before the system becomes unstable?
- h
6
k(s+1)
s2 +cs+1
esT
1
For this purpose, we plot the Nyquist and Bode digrams of the nominal transfer
function (s + 1)/(s2 + s + 1). The Matlab command margin gives numerical values
for the amplitude- and phase-margins.
Gm = Inf, Pm = 109.47 deg (at 1.4142 rad/sec)
Nyquist diagram
Magnitude
1.5
Im
0.5
0.5
0
0
45
0.5
Re
1.5
Phase
1.5
0
90
135
180
Frequency
10
Figure 2.4
12
Nyquist and bode plots for the nominal transfer function (s + 1)/(s2 + s + 1)
10
2.2
Sensitivity
s+1
esT
+s+1
For small values of T the Nyquist plot will not encircle 1, so the systems remains
stable. To find out exactly how large values of T are needed for instability, note that
the phase margin 109 degrees (or 109 /180 radians) is obtained at the frequency
1.41 rad/sec. Hence a time delay of
109
= 1.35 seconds
180 1.41
can be tolerated for k = c = 1.
To investigate the robustness to variations in the parameters c and k, we note
that the closed loop characteristic polynomial for T = 0 is
2.2 Sensitivity
Two transfer functions are of particular interest in the study of the feedback loop
below.
n
d
e
u
C ( s)
P ( s)
Figure 2.5
These are
1
1 + C ( s) P ( s)
C ( s) P ( s)
T ( s) =
1 + C ( s) P ( s)
S ( s) =
Lecture 2.
1
= sup
1 + C (i ) P(i )
Re
C (i ) P(i )
Figure 2.6 The L2 -gain of the sensitivity function is the inverse of the distance from the
Nyquist plot to 1.
the system, while C (s) P(s) is called the open loop transfer function or just the
loop transfer function.
The name sensitivity function refers to the fact that S measures how small
relative errors in P are mapped into relative errors in T. This is verified by a
simple calculation:
d
dT
=
dP
dP
1
1
1 + CP
C
TS
=
(1 + CP)2
P
dT / T
=S
dP/ P
e1
S1
e2
r2
S2
In this section, we will investigate stability robustness using the following
theorem, based on the notion of input-output L2 -gain. For simplicity, calculations
are done assuming zero initial conditions.
THEOREM 2.3THE SMALL GAIN THEOREM
Assume that S 1 and S 2 are input-output stable systems with L2 -gain qS 1 q and
qS 2 q. If qS 1 q qS 2 q < 1, then the L2 -gain from (r1 , r2 ) to ( e1 , e2 ) in the closed loop
14
2.3
system
e1 = S 2 ( e2 ) + r1
e2 = S 1 ( e1 ) + r2
is finite.
Proof. Define q yqT =
qR
T
0
e1 = r1 + S 2 (r2 + S 1 ( e1 ))
q e1 qT qr1 qT + qS 2 q qr2 qT + qS 1 q q e1 qT
q e1 qT
q r 1 q T + qS 2 q q r 2 q T
1 qS 1 q qS 2 q
C ( s)
(s)
-?
f
P ( s)
Figure 2.7
The transfer function from w to v is equal to T (s), the complementary sensitivity function. Hence, by the small gain theorem, the feedback system remains
stable as long as
qq qT q < 1
Note that the small gain theorem does not assume linearity or time-invariance.
Hence the closed loop system will remain stable for all plants of the form P(s)[1 +
(s)] where has L2 -gain smaller than [sup pT (i )p]1 , even for that are
nonlinear or time-varying.
As a second example, let us derive a stability criterion for the case that the
perturbation appears additively, i.e. P(s) is replaced by P(s) + (s).
Then the transfer function from w to v is equal to C (s) S(s), so the small gain
theorem shows stability for all perturbations satisfying
qq q CSq < 1
For linear time-invariant perturbations, this criterion can be nicely illustrated in
the Nyquist diagram. Clearly, the condition
p C p < p1 + PC p
guarantees that the Nyquist plot of ( P + ) C does not encircle 1.
15
Lecture 2.
C ( s)
(s)
P ( s)
-?
f
1
Figure 2.8
1 + PC
16
Lecture 3
3.1 Specifications
Recall from Lecture 1 the illustration of the design process shown in Figure 3.1.
While Lecture 2 was mainly concerned with analysis, we are now focusing on the
three neighboring blocks: Specification, Analysis and Synthesis.
Matematical model
and
specification
Idea/Purpose
Analysis
Experiment
Synthesis
Implementation
Figure 3.1
We will restrict attention to the following structure (Figure 3.2), with a scalar
transfer function for the plant. This setup was studied in the basic course and is
sufficient for many practical situations.
The controller consists of two transfer functions, the feedback part C (s) and
the feedforward part F (s). The control objective is to keep the process output x
close to the reference signal r, in spite of load disturbances d. The measurement
y is corrupted by noise n.
1 Written
17
Lecture 3.
d
e
F ( s)
u
C ( s)
P ( s)
1
Controller
Figure 3.2
Process
A Controller with two degrees of freedom
X ( s) =
The signals in the feedback loop are characterized by four transfer functions
(sometimes called The Gang of Four)
1
1 + P ( s) C ( s)
P ( s)
1 + P ( s) C ( s)
C ( s)
1 + P ( s) C ( s)
P ( s) C ( s)
1 + P ( s) C ( s)
In particular, we recognize the first one as the sensitivity function and the
last one as the complementary sensitivity.
The total system with a controller having two degrees of freedom is characterized by six transfer functions (The Gang of Six).
To fully understand the properties of the closed loop system, it is necessary to look
at all the transfer functions. It can be strongly misleading to only show properties
of a few input-output maps, for example a step response from reference signal to
process output. This is a common mistake in the literature.
The properties of the different transfer functions can be illustrated in several
ways, by time- or frequency-responses. For a particular example, we show below
first the six frequency response amplitudes, then the corresponding six step responses.
It is worthwhile to compare the frequency plots and the step responses and to
relate their shape to the specifications A-D:
18
3.1
PC F /(1 + PC )
0
10
10
P/(1 + PC )
PC /(1 + PC )
10
10
10
10
C F /(1 + PC )
10
10
Specifications
10
10
C /(1 + PC )
10
10
10
1
10
1/(1 + PC )
10
10
1
10
10
10
10
10
1
10
10
10
10
10
10
10
10
10
Figure 3.3 Frequency response amplitudes for P(s) = (s + 1)4 , C(s) = 0.775(s1 /2.05 + 1)
when F (s) is designed to give PCF /(1 + PC) = (0.5s + 1)4
PC F /(1 + PC )
P/(1 + PC )
PC /(1 + PC )
1.5
1.5
1.5
0.5
0.5
0.5
10
20
C F /(1 + PC )
30
10
20
C /(1 + PC )
30
1.5
1.5
1.5
0.5
0.5
0.5
10
20
30
10
20
30
10
20
30
10
20
30
1/(1 + PC )
Figure 3.4 Step responses for P(s) = (s + 1)4 , C(s) = 0.775(s1 /2.05 + 1) when F (s) is
designed to give PCF /(1 + PC) = (0.5s + 1)4
Disturbance rejection The two upper right plots show the effect of the disturbance d in process output x and input v respectively. The resulting process error
should not be too large and should settle to zero quickly enough. The control input
would cancel the disturbance exactly if the mid upper step response would be an
ideal step. In a short time-scale this is impossible, since the control input will not
change until the effect of the disturbance has appeared in the process output and
been available for measurement. However, slow disturbances should normally be
cancelled by u. Equivalently, the sensitivity function 1/(1 + PC ) should be small
for low frequencies. This specification is usually corresponds to an integrator in
the controller.
Supression of measurement noise The second specification was to limit the
effect of measurement noise, typically a high frequency phenomenon. The mid upper frequency plot shows good attenuation of measurement noise above the cut
off frequency of 1 Hz. In this example, this is mainly an effect of the process dynamics. A more interesting question is maybe the gain from measurement noise
to control input, since fast oscillations in the control actuator are usually undesir-
19
Lecture 3.
able. For this aspect, the mid lower frequency plot, showing the Bode amplitude
from n to v, is of interest.
Robustness to process variations As shown in the previous lecture, the robustness to process variations is determined by the sensitivity functions. In this
example, the lower right frequency plot has a maximal value of 2, which shows
that a small relative error in the process can give rise to a relative error of double
size in the closed loop transfer function. The maximal amplitude of the frequency
plot for the complementary sensitivity function is 1.35, so the small gain theorem
proves stability of the closed loop system as long as the relative error in the process model is below 74% = 1/1.35. In fact, most process models are inaccurate at
high frequencies, so the complementary sensitivity function PC /(1 + PC ) should
be small for high frequencies.
Command response The upper left corner plot shows the map from reference
signal r to process output x. Using the prefilter F, it is possible to get a better
step response here than in the upper mid plot. The prize to pay is that the corresponding response in the control signal gets higher amplitude. This can be seen
by comparing the lower left plot, showing the map from r to v, to the lower mid
plot, which shows the corresponding map when F " 1.
20
Phase
Magnitude
3.2
20
20
10
10
0
0
0
60
30
30
60
10
Figure 3.5
(right)
10
10
0 2
10
s+10
s+1
Loop shaping
10
10
10
10s+1
s+1
Lead compensator
Increases high frequency gain: Can be used for faster closed loop response
Increases phase, which may improve stability margins
Phase
Magnitude
Loop shaping design of high order controllers will be exercised in lab 1. We will
first design a controller C1 (s) for low frequencies, then keep adding compensator
links C2 (s), C3 (s), . . . to modify the dynamics at higher and higher frequencies
until a satisfactory controller C (s) = C1 (s) Cm (s) is obtained. Lead/lag links
are often sufficient, but occasionally it is useful to also consider controllers with
poles or zeros outside the real axis. The figure below shows the Bode diagram for
cases with stable complex zeros (left) and complex poles (right).
0
30
10
20
20
10
30
90
0
90
45
45
45
45
90
10
Figure 3.6
Notch compensator
10
s2 +0.1s+1
(s+1)2
90
10
10
(s+1)2
s2 +0.1s+1
(right)
21
Lecture 3.
Mu
My
um
ym
u
C
1
The two configurations are mathematically equivalent provided that
C F = Mu + CMy
The transfer functions Mu and My can be viewed as generators of the desired
output ym and the corresponding input um .
The transfer function from r to the error signal e = ym y is ( My PMu ) S.
The error is zero provided that
Mu = My / P
Notice that this condition does not depend on C! Since Mu = My / P should be
stable, causal and not include derivatives we find that
1
(s + 1)4
My (s) =
1
(sT + 1)4
then
Mu (s) =
(s + 1)4
My (s)
=
P ( s)
(sT + 1)4
Mu ()
1
= 4
Mu (0)
T
Fast response (T small) requires high gain of Mu . Bounds on the control signal
therefore limit how fast response we can obtain.
2
22
Lecture 4
Track
Disk
Pit
Sledge
Pickup head
trackpitch
Figure 4.1 Pits forming tracks on DVD surface (left). Larger radial movements are taken
care of by the sledge (right).
The surface velocity is constant (about 3.5 m/s), meaning that the disc should
rotate at different speeds depending on the current reading position.
23
Lecture 4.
represented by pits of different lengths in tracks on the disk. The image of the
surface is reflected back through the lens and read by a set of four photo detectors.
These pits make the laser beam interfere destructively with itself, and therefore
the pits look black to the detector.
Tracks
Radial electromagnet
A B
C D
Lens
A B
C D
Springs
Light detectors
Focus electromagnet
Laser
Figure 4.2 The pick-up-head has two electromagnets for fast positioning of the lens (left).
The four photo detectors A D (right).
Focus Error
Correct focus
Lens height
Too low
Figure 4.3
24
Too high
Sweeping the lens axially from low to high results in a curve like this.
4.1
Radial error
The radial error ( RE), meaning sideways deviation from the track center, can be
measured in two ways using the photo detectors. The simple way is to let
RE = A + C ( B + D ),
i.e., use the difference in light from the left and right pair of detectors. For example,
if the reflected light is brighter to the left, the radial error is positive, and we
should move right. This measurement method is called radial push-pull (PP). See
Figure 4.4.
A B
C D
Pit
Figure 4.4
There is a second measurement method (DPD), which usually gives better result,
but requires the disk to have pits. This is not true for a non-written DVD-R
(writable), for example. The signals f1 = A + D and f2 = B + C are created,
and phase compared (see Figure 4.5). For example, if f1 comes before f2 the lens
is too far to the right. The time difference forms the error signal RE.
A B
C D
f1
f2
Radial error
Pit
Figure 4.5
According to the DVD specification, DPD should be used whenever possible. Sweeping the lens radially over the disk creates an RE as in Figure 4.6 for PP and DPD.
1 track
DPD
PP
Radial position
Figure 4.6 DPD- and PP-signals as the disk rotates and the lens is swept over the tracks
radially. As can be seen, DPD is linear in a larger range.
25
Lecture 4.
4.2 Dynamics
The DVD player system can be viewed as a two-input, two-output dynamical system. The main dynamics are due to the springs and masses in the lens system.
The inputs are voltages to the electromagnets moving the lens, and the outputs
are voltages corresponding to F E and RE. See Figure 4.7.
ufocus
FE
Pick-up & Disc
uradial
RE
The pick-up head and disk seen as a two-input, two-output dynamical system.
Figure 4.7
The department has a raw DVD player without any controller. Using system
identification techniques, the transfer function P f (s) from ufocus to F E and the
transfer function Pr (s) from uradial to RE have been estimated. The cross-coupling
between inputs and outputs have been ignored for simplicity. The resulting Bode
diagrams can be seen in Figure 4.8.
Bode Diagram
Bode Diagram
100
Magnitude (dB)
50
50
50
50
90
90
Phase (deg)
Phase (deg)
Magnitude (dB)
100
180
270
2
10
10
10
Frequency (rad/sec)
10
10
180
270
1
10
10
10
Frequency (rad/sec)
10
10
Figure 4.8 Left: Transfer function estimate for the focus servo. The model is of second
order. Right: Transfer function estimate for the radial servo.
p C (i ) P(i )p 1000
for 23.1 Hz
p C (i ) P(i )p 1
The first specification is chosen to remove enough of the disk oscillation disturbances to stay in track. The second specification is chosen to reduce the effects of
measurement noise and small disturbances (dirt and scratches).
26
4.4
Focus controller
ufocus
uradial
FE
PUH & Disk
RE
Radial controller
Figure 4.9
Meeting these specifications is not a trivial task. In fact, most DVD readers
probably dont! Instead, the manufacturers modify the specifications according
to the circumstances. For example, a CD player for a car would need very good
disturbance rejection (high gain at low frequencies) and this is obtained at the
expense of high gain also at high frequncies, which gives less robustness to disc
scratches.
The goal of this lecture is to design the focus controller according to the specifications and try it on the experimental setup. In Lab 3 in this course, you will
design your own radial controller to keep the DVD player in track!
Figure 4.10
It is natural to introduce lag compensation to increase the gain at low frequencies. However, the break points need to be at frequencies well below 2 kHz
in order to avoid additional phase lag at the cut-off frequency. Using a lag filter
C1 (s) = 0.4 s+s600 gives the modified plots in Figure 4.11 (left).
27
Lecture 4.
Figure 4.11 A lag filter has been used to increase the gain at low frequencies (left). A lead
filter improves stability by increasing the phase near 2 kHz
Figure 4.12 Bode and Nichols plots with three lag filters and one lead filter (dashed) and
without compensation (solid).
28
Lecture 10
Fundamental limitations1
10.1 Introduction
Maybe the cardinal mistake in control engineering would be to consider the process
as fixed once and for all. In fact, the control system specifications could very well
be impossible to meet once the process is constructed and fixed. A striking example
of this is given in the following citation from F. R. Whitt and D. G. Wilson (MIT
Press, 1974), Bicycling Science - Ergonomics and Mechanics:
Many people have seen theoretical advantages in the fact that front-drive,
rear-steered recumbent bicycles would have simpler transmissions than
rear-driven recumbents and could have the center of mass nearer the front
wheel than the rear. The U.S. Department of Transportation commissioned
the construction of a safe motorcycle with this configuration. It turned out
to be safe in an unexpected way: No one could ride it.
This lecture is devoted to the fundamental limitations that are inherited from
properties of the controlled plant and will address questions like the following
two:
y(t) e zt dt
29
Lecture 10.
Fundamental limitations1
Figure 10.1 Schematic picture of a bicycle. The top view is shown on the left and the rear
view on the right.
Clearly the integral cannot be zero unless y(t) takes both positive and negative
values. The time duration of such dynamics is approximately 1/ z and limits the
achievable rate of control. Hence, while an unstable pole requires a fast feedback
loop, an unstable zero gives an upper bound on how fast it can be. A combination
of the two phenomena can make the system impossible to control.
Example 1 A tourque balance for a bicycle can be written as
mV0 {
d
d2
V0 + a
J 2 = m{ +
dt
b
dt
where the physical parameters have typical values as follows:
Mass:
m = 70 kg
Distance rear-to-center:
a = 0.3m
{ = 1.2 m
b = 0.7 m
Moment of inertia:
J = 120 kgm2
V0 = 5 ms1
Speed:
Acceleration of gravity:
= 9.81 ms2
mV0 { as + V0
b Js2 m{
30
10.2
Riding the bicycle at this speed, the zero is not really an obstacle for control.
However, with a rear-wheel-steered bicycle, the speed gets a negative sign and
the zero becomes unstable. In particular, for slow speed (( 0.7m/s) there is an
unstable pole-zero cancellation, which is impossible to stabilize.
2
p zp=1
1+z
1z
is analytic in the unit disc. Hence the Maximum Modulus Theorem can be applied
to give the following corollary:
COROLLARY 10.1
Suppose that all poles of the rational function G (s) have negative real part. Then
max p G (s)p = max p G (i )p
R
Re s0
S( z) :=
1
=1
1 + C ( z) P( z)
Notice that the unstable zero in the plant can not be cancelled by an unstable pole
in the controller, since this would give an unstable transfer function C /(1 + CP)
from measurement noise to control input.
Similarly, the complimentary sensitivity must be one at an unstable pole:
P( p) =
T ( p) :=
C ( p) P( p)
=1
1 + C ( p) P( p)
31
Lecture 10.
Fundamental limitations1
log p Wa p
log p W b p
0
a
log
log
Figure 10.2 Amplitude plots for weighting functions. The left weighting function is used
to bound the sensitivity at small frequencies, while the right function is used to boud the
complementary sensitivity at high frequencies
q S q 1
This bound is however not particularly interesting, since usually S(i ) ( 1 for high
frequencies anyway. A much more interesting conclusion will next be obtained by
using a weighting function.
Recall that disturbance rejection requires small sensitivity for small frequencies. One way to formalize this condition is to define
Wa (s) =
s+a
2s
(10.1)
for some value of a. See Figure 10.2, left. Satisfying (10.1) with a high value of a
means fast disturbance rejection.
The specification requires that S(s) has a zero in the origin. This is often
obtained by an integrator in the controller. Moreover, Corollary 10.1 implies that
sup p Wa (i ) S(i )p = sup p Wa (s) S(s)p p Wa ( zi )p
Re s0
for every unstable zero zi of the plant P. In particular, the specification (10.1)
is impossible to satisfy unless p Wa ( zi )p 1, or in other words a zi , for every
unstable zero zi . Hence the unstable zeros give an upper bound on the achievable
bandwidth. In the following theorem, this discussion is summarized together with
a corresponding argument for unstable poles:
THEOREM 10.2
Suppose that the plant P(s) has unstable zeros zi and unstable poles p j . Define
the weighting functions Wa = (s + a)/(2s) and W b (s) = (s + b)/(2b). Then the
specifications
sup p Wa (i ) S(i )p 1
32
sup W b (i )T (i ) 1
10.4
b max p j
j
Proof. The statement about the sensitivity function was proved above, and the
statment about the complementary sensitivity function is analogous.
2
Example 2 Let us see what Theorem 10.2 has to say
p about the bicycle example.
The unstable pole gives a bound q W b T q 1 for b m{/ J. This shows that the
closed loop transfer function from measurement
noise to process output can not
p
be forced small for frequencies below m{/ J. A loose interpretation is that it
is impossible to ride the bicycle and keep the eyes shut except for a sample every
second. This applies for the bicycle with normal steering regardless of speed.
For a rear wheel steering bike, there is the second complication of an unstable
zero at V0 / a, which gives a bound on how fast disturbances one can reject. For
low speed, only slow disturbances can be rejected.
The special difficulties corresponding to a combination of an unstable pole and
an unstable zero nearby are however not apparent in Theorem 10.2. Such problems
will be treated next.
2
The following theorem gives simple expressions for the limitations caused by
an unstable pole/zero pair.
THEOREM 10.3
If P(s) has an unstable pole p and an unstable zero z, then
1
z + p
z p
1 + CP
Note that if S is very large, then the same is true for T, since S + T " 1. Hence, if
p( z + p)/( z p)p is significantly larger than one, the system is impossible to control
because of poor robustness to model errors and amplification of measurement
noise.
b(s), with P
b proper and P
b( p) ,= 0. Then
Proof. Assume that P(s) = (s z)(s p)1 P
the sensitivity function satisfies
1
1
= sup
q Sq = sup
b(i z)(i p)1
1 + CP
1 + CP
i p
i + p
= sup
= sup
b(i z)
b(i z)
i p + C P
i p + C P
z + p
s+ p
= sup
b
z p
Re s0 s p + C P( s z)
p
The fourth inequality uses that pi pp = 2 + p2 = pi + pp and the fifth
inequality is Corollary 10.1.
2
A similar argument can be applied to a system involving a time delay but
application of the maximum modulus theorem is less straightforward in this case.
33
Lecture 10.
Fundamental limitations1
0.1
magnitude
10
10
0.1
10
0.2
10
0.3
10
frequency
Figure 10.3 The amplitude curve of the sensitivity function always enclose the same area
below the level p Sp = 1 as above.
THEOREM 10.4
If P(s), C (s) and S(s) = [1 + C (s) P(s)]1 are stable and s2 C (s) P(s) is bounded,
then
Z
log p S(i )p d = 0
0
Proof. Proof sketch. From the theory of analytic functions, recall that Cauchys
formula states that
Z
f ( z)dz = 0
for every closed path in the region where the function f is analytic. Bodes
integral formula follows by application of Cauchys formula to
f ( z) = log S( z)
The stability of C and P guarantee that f is well-defined and analytic in the
whole right half plane. Integration along the imaginary axis can be extended to
integration along a closed path by adding a large half-circle in the right half
plane. The condition that s2 C (s) P(s) is bounded is needed to make sure that the
contribution from the half-circle vanishes as the radius tends to infinity.
2
The invariance of Bodes integral is sometimes referred to as the water-bed
effect: If the designer tries to push the magnitude of the sensitivity function down
at some point, it will inevitably pop up somewhere else!
The assumptions behind Bodes integral formula deserve some discussion. The
expression s2 C (s) P(s) is always bounded whenever C (s) and P(s) correspond to
real sensor/actuator interconnections, since direct terms are not physically implementable. With unstable poles in C (s) P(s) the integral formula changes into
log p S(i )p d =
Re pi
which makes it even harder to push down the sensitivty magnitude! The faster
unstable modes, the harder it is. In fact, this can be used as an argument why
unstable controllers should in general be avoided.
34
Lecture 11
Multivariable control1
There is a clear trend in modern engineering toward systems of higher and higher
complexity. One reason is that demands for efficiency give tighter interconnections
between subsystems. For example, to get minimally pollutive emissions in the
exhaust gas of a car, it is necessary that engine, carburettor, catalyst, gearbox, etc.
all cooperate in an optimal manner. Similarly, the demand for efficient production
and distribution of electrical power has led to tighter coupling between production
units in different geographical regions and more complex large scale dynamics.
A system with several inputs and several outputs is sometimes called a MIMO
(Multiple-Input-Multiple-Output) system. Control theory for such systems is a
highly active research area and a review of the available methods is outside the
scope of this course. However, many of the ideas that were developed for scalar
systems can be easily adapted also to a multivariable setting. This lecture will
present a few such items:
Figure 11.1 A modern car, a power plant and an oil refinery all make extensive use of
multivariable control systems
35
Lecture 11.
Multivariable control1
n
d
r
e
F ( s)
C ( s)
P ( s)
I
Controller
Process
Figure 11.2
[ I + PC ]1 P = P[ I + CP]1
C [ I + PC ]1 = [ I + CP]1 C
T = P[ I + CP]1 C = PC [ I + PC ]1
S+T = I
The first equality follows by multiplication with I + CP from the right and I + PC
from the left. The second one is analogous. Using the first two equalities, we
immediately get the third. The last one is straight from definitions as well.
Also for multivariable systems, it is common to require S to be small at low
frequencies and T to be small at high frequencies. The first specification means
that y follows Fr well at small frequencies, while the second means that high
frequency measurement noise n does not influence x significantly. Another way
to state these requirements is to say that the loop transfer matrix
P(i ) C (i )
should have small norm q P(i ) C (i )q at high frequencies, while at low the frequencies instead q[ P(i ) C (i )]1 q should be small. See Figure 11.3.
q WS Sq 1
is impossible to satisfy unless q WS ( z)q 1 for every unstable zero z of P(s).
36
11.2
Magnitude
20
P(i ) C (i )
10
0
10
20
30
Robustness
40
50
60
w0
70
w1
10
Frequency
0
10
Disturbance rejection
Figure 11.3 Specifications on the singular values of the loop transfer function often have
this form. A lower bound on the singular values for low frequencies is needed for disturbance
rejection. This means that q[ P(i ) C(i )]1 q should be small. An upper bound on the singular
values for high frequencies, making q P(i ) C(i )q small enough, is needed for robustness to
model errors and measurement noise.
The proof is analogous to the result for scalar systems in Lecture 4. Instead of
giving the details, we turn our attention to an example.
Example 1 (Non-minimum phase MIMO System) Consider a feedback system
y = ( I + PC )1 r with the multivariable process
2
s+1
P ( s) =
1
s+1
3
s+2
1
s+1
2
3
s + 1
=
(s + 1)2
(s + 2)(s + 1)
(s + 1)2 (s + 2)
shows that the process has an unstable zero at s = 1, which will limit the achievable performance. For further understanding of the limitation, consider the following three different control structures:
Controller 1
The controller
K 1 (s + 1)
s
C1 (s) =
K 1 (s + 1)
3K 2 (s + 0.5)
s(s + 2)
2K 2 (s + 0.5)
s(s + 1)
K 1 (s + 1)
s(s + 2)
P(s) C1 (s) =
K 2 (s + 0.5)(s + 1)
s(s + 1)(s + 2)
Hence the system is decoupled into to scalar loops, each with an unstable zero
at s = 1 that limits the bandwidth. The closed loop step responses are shown in
Figure 11.4.
37
Lecture 11.
Multivariable control1
Step Response
1.5
Amplitude
1
0.5
0
0.5
5
Time (sec)
10
10
Step Response
1.5
Amplitude
1
0.5
0
0.5
5
Time (sec)
Figure 11.4 Closed loop step responses with decoupling controller C1 (s) for the two outputs
y1 (solid) and y2 (dashed). The upper plot is for a reference step for y1 . The lower plot is for
a reference step for y2 .
Controller 2
The controller
K (s + 1)
1
K2
s
C2 (s) =
K 1 (s + 1)
K2
s
gives the upper triangular loop transfer matrix
K 1 (s + 1)
K 2 (5s + 7)
s(s + 2)
(s + 2)(s + 1)
P(s) C2 (s) =
2K 2
0
s+1
The controller
C3 (s) =
K1
K1
K 2 (s + 0.5)
s(s + 2)
2K 2 (s + 0.5)
s(s + 1)
K 1 (5s + 7)
0
(s + 1)(s + 2)
P(s) C3 (s) =
K 2 (1 + s)(s + 0.5)
2K 1
s+1
s(s + 1)2 (s + 2)
In this case y1 is decoupled from r2 and can respond arbitrarily fast for high values
of K 1 , at the expense of bad behavior in y2 . Step responses for K 1 = 10, K 2 = 1
are shown in Figure 11.6.
To summarize, the example shows that even though a multivariable unstable
zero always gives a performance limitation, it is possible to influence where the
effects should show up.
2
38
11.3
Pairing of signals
Step Response
1.5
Amplitude
1
0.5
0
0.5
5
Time (sec)
10
10
Step Response
3
Amplitude
2.5
2
1.5
1
0.5
0
0.5
5
Time (sec)
Figure 11.5 Closed loop step responses with controller C2 (s) for the two outputs y1 (solid)
and y2 (dashed). The right half plane zero does not prevent a fast y2 -response to r 2 but at the
price of a simultaneous undesired response in y1 .
Step Response
1.5
Amplitude
1
0.5
0
0.5
5
Time (sec)
10
10
Step Response
Amplitude
0.5
0.5
5
Time (sec)
Figure 11.6 Closed loop step responses with controller C3 (s) for the two outputs y1 (solid)
and y2 (dashed). The right half plane zero does not prevent a fast y1 -response to r 1 but at the
price of a simultaneous undesired response in y2 .
39
Lecture 11.
Multivariable control1
Inputs:
40
4
e27s
Y1 (s)
50s + 1
=
5.4
Y2 (s)
e18s
50s
+1
|
1.8
e28s
60s + 1
5.7
e14s
60s +{z1
P ( s)
5.9
U1 ( s )
e27s
50s + 1
U2 ( s )
6.9
e15s
U3 ( s )
40s + 1
}
11.3
Figure 11.7
Pairing of signals
0.6111 1.3285
RGA( P(0)) =
0.0134 1.5827 0.5962
0.4355 0.3667i 0.6536 0.0171i 1.2181 + 0.3839i
RGA( P(i/50)) =
0.0906 + 0.3667i 1.5933 + 0.0171i 0.5027 0.3839i
0.2827
To choose control signal for y1 , we apply the rules of thumb to the top row. This
suggests the bottom temperature u3 for control of the top draw composition y1 ,
since the third column has values slightly closer to 1 than the first column.
Based on the bottom row, we choose the side draw flowrate u2 to control the
side draw composition y2 . The top draw flow rate u1 is left unused. The matrix
transfer function P (s) from (u3 , u2 ) to ( y1 , y2 ) is now to be controlled by a diagonal
controller C (s), say a PI controller:
60s + 1
5.9e27s 1.8e28s
0
50s + 1 60s + 1
C (s) = 50s
P (s) =
6.9e15s 5.7e14s
60s + 1
0
50s
40s + 1 60s + 1
Without feedforward, the closed loop transfer matrix from reference to output
becomes
PC ( I + PC )1
41
Lecture 11.
Multivariable control1
From: U(1)
From: U(2)
20
To: Y(1)
0
20
40
Magnitude (dB)
60
80
20
To: Y(2)
0
20
40
60 2
10
10
2 2
1010
10
10
Frequency (rad/sec)
Figure 11.8
Magnitude plots for the closed loop transfer function of the distillation column
and the Bode magnitude plots for the four transfer functions are given in Figure 11.8. As seen in the plots, the resulting cross-coupling is generally small and
there is no static error.
42