Astrom Libro
Astrom Libro
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2 A Brief History . . . . . . . . . . . . . . . . . . . . . . . . 10
1.3 Process Control . . . . . . . . . . . . . . . . . . . . . . . . 14
1.4 Manufacturing . . . . . . . . . . . . . . . . . . . . . . . . 17
1.5 Robotics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
1.6 Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.7 Aeronautics . . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.8 Electronics and Communication . . . . . . . . . . . . . . 28
1.9 Automotive . . . . . . . . . . . . . . . . . . . . . . . . . . 33
1.10 Computing . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
1.11 Mathematics . . . . . . . . . . . . . . . . . . . . . . . . . 39
1.12 Physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
1.13 Biology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
1.14 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
2. Feedback . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 45
2.2 Simple Forms of Feedback . . . . . . . . . . . . . . . . . 46
2.3 Representation of Feedback Systems . . . . . . . . . . . 49
2.4 Properties of Feedback . . . . . . . . . . . . . . . . . . . 58
2.5 Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
2.6 Open and Closed Loop Systems . . . . . . . . . . . . . . 66
2.7 Feedforward . . . . . . . . . . . . . . . . . . . . . . . . . . 68
2.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
3. Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.2 Two Views on Dynamics . . . . . . . . . . . . . . . . . . . 72
3.3 Linear Differential Equations . . . . . . . . . . . . . . . 77
3.4 Laplace Transforms . . . . . . . . . . . . . . . . . . . . . 86
3
Contents
4
Contents
5
Contents
6
1
Introduction
1.1 Introduction
To do
• Clean up what is written
– Wal-Mart
– Lamp
– Sled in CD picture
• Decide on figures and make drafts
– Annan bild pa centrifugal governor som visar angventilen
– Better picture of adaptive optics
– Combine Fig 1.2 and 1.3 to one figure
– Block diagram for step length control
– Human posture
– Human posture
– Aibo mm Sony robota
• Organize and add
• References
Control systems are ubiquitous. They appear in our homes, in cars, in
industry and in systems for communication and transport, just to give a
few examples. Control is increasingly becoming mission critical, processes
will fail if the control does not work. Control has been important for design
of experimental equipment and instrumentation used in basic sciences
7
Chapter 1. Introduction
and will be even more so in the future. Principles of control also have an
impact on such diverse fields as economics, biology, and medicine.
Control, like many other branches of engineering science, has devel-
oped in the same pattern as natural science. Although there are strong
similarities between natural science and engineering science it is impor-
tant to realize that there are some fundamental differences. The inspi-
ration for natural science is to understand phenomena in nature. This
has led to a strong emphasis on analysis and isolation of specific phe-
nomena, so called reductionism. A key goal of natural science is to find
basic laws that describe nature. The inspiration of engineering science
is to understand, invent, and design man-made technical systems. This
places much more emphasis on interaction and design. Interaction is a
key feature of practically all man made systems. It is therefore essential
to replace reductionism with a holistic systems approach. The technical
systems are now becoming so complex that they pose challenges compara-
ble to the natural systems. A fundamental goal of engineering science is
to find system principles that make it possible to effectively design com-
plex systems. Feedback, which is at the heart of automatic control, is an
example of such a principle.
A simple form of feedback consists of two dynamical systems connected
in a closed loop which creates an interaction between the systems. Simple
causal reasoning about such a system is difficult because, the first system
influences the second and the second system influences the first, leading to
a circular argument. This makes reasoning based on cause and effect diffi-
cult and it is necessary to analyze the system as a whole. A consequence of
this is that the behavior of a feedback system is often counterintuitive. To
understand feedback systems it is therefore necessary to resort to formal
methods.
Feedback has many advantages. It is possible to create linear behavior
out of nonlinear components. Feedback can make a system very resilient
towards external influences. The total system can be made very insensi-
tive to external disturbances and to variations in its individual compo-
nents. Feedback has one major disadvantage, it may create instability,
which is intrinsically a dynamic phenomenon. To understand feedback
systems it is therefore necessary to have a good insight into dynamics.
The wide applicability of control has many advantages. Since control
can be used in so many different fields, it is a very good vehicle for tech-
nology transfer. Ideas invented in one field can be applied to another
technical field.
Control is inherently multi-disciplinary. A typical control system con-
tains sensors, actuators, computers and software. Analysis of design of
control systems require knowledge about the particular process to be con-
trolled, knowledge of the techniques of control and specific technology
8
1.1 Introduction
9
Chapter 1. Introduction
Although there are early examples of the use of feedback in ancient his-
tory, the development of automatic control is strongly connected to the
industrial revolution and the development of modern technology. When
new sources of power were discovered the need to control them immedi-
ately arose. When new production techniques were developed there were
needs to keep them operating smoothly with high quality.
10
1.2 A Brief History
Figure 1.1 The centrifugal governor, which has been used to control the speed of
engines since the beginning of the industrial revolution. When the axis spins faster
the balls move away from the axis. The motion is transfered to the engine to reduce
its power. The governor has also become an icon of the field of control.
system is a feedback system because changes in the velocity are fed back
to the steam valve. The feedback is negative because the the steam supply
is increased when the velocity decreases.
The improvement obtained when using a centrifugal governor is illus-
trated in Figure 1.2. The figure shows that the velocity drops when an
additional loom is connected to the drive belt. The figure also shows that
the velocity drop is significantly smaller when a centrifugal governor is
used.
It is possible to change the characteristics of the governor by changing
the mechanism that transmits the motion of the balls to the steam valve.
To describe this we introduce the notion of gain of the governor. This is
the ratio of the change in steam valve opening (∆ u) to the change (∆ v)
of the velocity. See Figure 1.2 which shows how the velocity responds to
changes in the load. The figure shows that the velocity error decreases
with decreasing gain of the controller but also that there is a tendency
for oscillations. The system becomes unstable for large gain. The centrifu-
gal governor was a very successful device that drastically simplified the
operation of steam driven textile mills.
The action of the basic centrifugal governor can crudely be describe
with the equation
u = k( Vr − V ) + b
11
Chapter 1. Introduction
Velocity
Velocity
2 2
1 1
0 0
0 2 4 6 8 10 0 2 4 6 8 10
4 4
PSfrag replacements 3 3
Load
Load
2 2
1 1
0 0
0 2 4 6 8 10 0 2 4 6 8 10
Time Time
Figure 1.2 Response of the velocity to changes in the load of an engine controlled
by a governor with proportional control left and proportional and integral control
right. The values of the gain in the left figures are, k = 0 (dashed), k = 1 (full),
k = 12 (dash-dotted). In the right figure the dashed curve represents proportional
control ( k = 1) and the full line PI control ( k = 1 and ki = 0.5).
the bias term was thus replaced by a term proportional to the integral of
past errors. A controller described by (1.1) has the amazing property that
the velocity V is always equal to the desired velocity V r in steady state.
This is illustrated by Figure 1.2 which shows the behavior of an engine
with control actions proportional to the integral of the error. In standard
terminology the Siemens governor is called a PI controller, indicating that
the control action is proportional to the error and the integral of the error.
Integral action has some amazing properties that will be discussed further
in Section 2.2.
12
1.2 A Brief History
13
Chapter 1. Introduction
from mathematics. The major drivers for the development were the space
race and the use of digital computers for industrial process control. Notice
that Sputnik was launched in 1957 and the first procss control computer
was installed in an oil refinery in 1959. There was a very vigorous devel-
opment of both theory and practice of control. Digital computers replaced
analog computers both for simulation and implementation of controllers.
A number of subspecialities of control were also developed. It turned out
that a wide range of mathematics fitted the control problems very well,
and a tradition of a high respect for mathematical rigor emerged.
Today control is a well established field with a solid body of theory
and very wide applications. Practically all controllers are today imple-
mented in digital computers. There are also many challenges. Control is
increasingly becoming mission critical which requires that safety and re-
liability are very important. The complexity of the control systems are
also increasing substantially.
There are tremendous challenges in the future when we can visualize a
convergence of control, computing and communication which will result in
large interconnected systems. It is also reasonable to guess that a deeper
understanding of control in the field of biology will be very exciting. In
this sections we will present a number of applications together with some
historical notes. The idea is to illustrates the broad applicability of control
and the way it has impacted on many different fields.
Many products that are essential for us in daily life are manufactured
by the process industries. Typical examples are oil, petrochemical, pa-
per, pharmaceuticals. These industries use continuous manufacturing pro-
cesses. Process control is an essential part of these industries, the pro-
cesses can not be run without control systems. Control of engine speed
using the centrifugal governor represent early examples of process control,
see Section 1.2.
14
1.3 Process Control
Figure 1.3 Microprocessor based single loop controller. By courtesy of ABB In-
dustrial Systems.
15
Chapter 1. Introduction
Plant
manager
Lab Process
assistant Plant network operator
Control network
Process
Figure 1.4 Modern industrial systems for process control, like the Advant OCS
tie computers together and help create a common uniform computer environment
supporting all industrial activities, from input to output, from top to bottom. (By
courtesy of ABB Industrial System, Västerås, Sweden.).
16
1.4 Manufacturing
1.4 Manufacturing
17
Chapter 1. Introduction
Supply Chains
There are also other uses of control in manufacturing, namely control of
the complete business operation and complete supply chains. Manufac-
turing and distribution involve large flows of materials. Raw materials
have to be supplied to the manufacturer, the products have to be supplied
to the consumers. This is illustrated in Figure 1.5 which shows a simple
supply chain. The diagram shows only a few levels, in practice a system
may contain thousands of outlets and storages and it may deal with a
very large number of products. Several important functions in the system
like quality control, sales, and production management are not shown in
the figure.
Manufacturing facilities and sales operations have traditionally been
quite inflexible in the sense that it is difficult to change products and
production rates. In such a system it is necessary to have many buffer
storages to match production rates to fluctuating sales and supplies of
raw materials and parts from sub-contractors. Production rate is largely
determined in an open loop manner based on prediction of sales. An elab-
orate bookkeeping system is also required to keep inventories of parts.
A different system is obtained if the production time can be reduced
and if the production can be made so flexible that products can be changed
rapidly. It is then possible to produce a part when an order is obtained.
Such a system may be regarded as a closed loop system where production
is determined by feedback from sales. An advantage of such a system is
that inventories can be reduced significantly. Administration of the system
is also greatly simplified because it is no longer necessary to keep track
of many products and parts in storage. A system of this type contributed
significantly to the efficiency of Wal-Mart. Kolla They relied on human
sensors in the form of managers who analyzed the sales when the shops
were closed every evening to make new orders.
18
1.5 Robotics
Figure 1.6 Remote robot surgery using the ZEUS system from Computer Motion
Inc. The doctors on the left are in New York and the patient and the robots are in
Strasbourg, France. Courtesy of Computer Motion Inc. Goleta
1.5 Robotics
The origin of the industrial robot is a patent application for a device called
Programmed Article Transfer submitted in 1956 by the engineer George
Devol. The robotics industry was created when Devol met Joseph En-
gelberger and they founded the company Unimation. The first industrial
robot, Unimate, was developed by in 1961. A major breakthrough occurred
in 1964 when General Motors ordered 66 machines Unimate robots from
Unimation. In 1998 there were about 720 000 robots installed. The ma-
jority of them, 412 000 are in Japan. Robots are used for a wide range
of tasks: welding, painting, grinding, assembly and transfer of parts in
a production line or between production lines. Robots that are used ex-
tensively in manufacturing of cars, and electronics. There are emerging
applications in the food industry and in packaging. Robots for vacuum-
ing and lawn moving as well as more advanced service robots are also
appearing.
Robots are also started to be used used in medical applications. This
is illustrated in Figure 1.6 which shows robots that are used for an endo-
scopic operation. One advantage of the system is that it permits doctors
to work much more ergonomically. Another advantage is that operations
can be done remotely. The system in the figure is from a robot operation,
where the patient was in Strasbourg, France and the doctors in New York.
19
Chapter 1. Introduction
by the check author Karel Capek called Rossum’s Universal Robots. The
play is about robot workers who revolt and kill their human ruler. Another
literary aspect of robot is the famous robot trilogy by Isaac Asimov from
1940 who introduced the three laws of robotics
First Law: A robot may not injure a human being, or, through inaction,
allow a human being to come to harm.
Second Law: A robot must obey orders given it by human beings, except
where such orders would conflict with the First Law.
Third Law: A robot must protect its own existence as long as such pro-
tection does not conflict with the First or Second Law.
Asimovs book captured the imagination of many and there is clear evi-
dence that Engelberger was inspired by Asimovs writing. Other examples
about imaginative speculations about robots are found in Arthur Clarkes
book 2001 A Space Odyssey where the robot HAL takes over the operation
of a space ship and R2- D2 in the Star Wars series.
There are currently some very interesting developments in robotics.
Particularly in Japan there is much research on humanoid and animaloid
robots. There are very advanced humanoid robots in research laboratories,
there are also robots that mimic snakes, cats, birds and fish. A robot dog
AIBO and a service robot have been commercialized by Sony.
20
1.6 Power
1.6 Power
The power industry started to develop in the end of the 19the century and
accelerated rapidly in the 20th century. Availability of power has improved
quality of life tremendously. In 2000 the total amount of electric energy
generated in the world was about about 15 000 TWh. It is expected to
grow by a factor of 3 in 60 years where the main growth is outside the
OECD countries.
Control is an essential element in all systems for generation and trans-
mission of electricity. Many central ideas in control were developed in this
context. When electricity emerged in the end of the 19th century the gen-
erators were typically driven by water turbines. Since alternating current
(AC) was the preferred means for transmission there was immediately a
need to control the speed of the generators to maintain constant frequency.
Derivative and integral control appeared early as did stability critiera.
The development paralleled the work on centrifugal governors but it was
done independently and it had a stronger engineering flavor. One of the
earliest books on control, with the title Die Regelung der Kraftmaschinen,
was published by Tolle as early as 1905. It was discovered that the perfor-
mance of hydroelectric power stations was severely limited by dynamics
of the water duct. The power decreased rapidly initially when the valve
was opened and then increased slowly. This property made the systems
difficult to control. It is an example of what is now call a non-minimum
phase dynamics.
As the demand for electricity grew many generators were connected in
a network. These networks became larger and larger as more generators
and consumers were connected. Figure 1.7, which is a schematic picture
of the network for the Scandinavian countries, is an example of a network
of moderate size. Sweden, Norway and Finland has much hydroelectric
power, Sweden and Finland has nuclear power and Denmark has wind
and thermal power. In Sweden the hydroelectric power is generated in
the north, but most of the consumption is in the south. Power thus has to
be transmitted over long lines. The system in the different countries are
connected via AC and DC lines. There are also connections to Germany
and Poland.
It is difficult to store energy and it is therefore necessary that pro-
duction and consumption are well balanced. This is a difficult problem
because consumption can change rapidly in a way that is difficult to pre-
dict. Generators for AC can only deliver power if the generators are syn-
chronized to the voltage variations in the network. This means that the
rotors of all generators in a network must line up. Synchronism is lost
when the angles deviate too much.
Matching production and consumption is a simple regulation problem
21
Chapter 1. Introduction
Figure 1.7 The Nordel power grid which supplies electric energy for the Scan-
dinavian countries. The squares represent hydroelectric stations and the triangles
represent thermal stations, both nuclear and conventional, and circles denote trans-
formers. The lines denote major power lines. Only the major components are shown
in the system.
22
1.6 Power
DC line
PSfrag replacements
Rectifier Inverter
Figure 1.8 Schematic diagram of an HVDC transmission link. The system is fed
from the right by AC which is converted to DC by the rectifier and transmitted over
a DC line to the inverter which converts it to AC.
for one generator and one consumer, but it is a more difficult problem
in a highly distributed system with long distances between consumption
and generation. To have a reliable system it is highly desirable to avoid
transmission of information over long distances. Control should therefore
be done locally at each station based on the information available at the
station. Several interesting control principles have been developed to do
this. Control of each generator must be based on information that is locally
available. Because of reliability requirements it is not possible to rely on
information that is transmitted over wide distances.
23
rag replacements
Rectifier
Inverter
DC line
Chapter 1. Introduction
1
10
0
10
-1
10
-2
10
10 4 10 5 10 6 10 7
Figure 1.9 Power outages in the US 1984-97. The horizontal axis shows the num-
ber of persons N affected by the outages and the vertical axis shows the yearly
frequency of outages that influence more than N persons. Notice that the scales on
both axes are logarithmic.
first observations that problems may occur when several regulators are
connected to an integrated system.
Edisons observation led to interesting developments of control theory.
In current practice one large generator in the network controls the fre-
quency using a controller with integral action. The other generators use
proportional control. The amount of power delivered by each generator is
set by the gain of the proportional controller. Each generator has separate
voltage control.
There have been many other surprises in interconnected systems. In
the Nordel system it has been observed that a moderated increase of
power load in the north could result in large oscillations in the power
transmission between Sweden and Denmark in the south. Oscillations
have been observed when modern trains with switched power electronics
have put in operation. An understanding of such phenomena and solutions
require knowledge about dynamics and control.
24
1.7 Aeronautics
1.7 Aeronautics
Control has often emerged jointly with new technology. It has often been
an enabler but in some cases it has had a much more profound impact.
This has been the case in aeronautics and astronautics as will be discussed
in this section.
Emergence of Flight
The fact that the ideas of control has contribute to development of new
technology is very nicely illustrated by the following quote from a lecture
by Wilbur Wright to the Western Society of Engineers in 1901:
“ Men already know how to construct wings or airplanes, which
when driven through the air at sufficient speed, will not only
sustain the weight of the wings themselves, but also that of
the engine, and of the engineer as well. Men also know how to
build engines and screws of sufficient lightness and power to
drive these planes at sustaining speed ... Inability to balance
and steer still confronts students of the flying problem. ... When
this one feature has been worked out, the age of flying will have
arrived, for all other difficulties are of minor importance.”
The Wright brothers thus realized that control was a key issue to enable
flight. They resolved compromise between stability and maneuverability
by building an airplane, Kitty Hawk, that was unstable but maneuverable.
The pioneering flight was in 1905. Kitty Hawk had a rudder in the front of
the airplane, which made the plane very maneuverable. A disadvantage
was the necessity for the pilot to keep adjusting the rudder to fly the
plane. If the pilot let go of the stick the plane would crash. Other early
aviators tried to build stable airplanes. These would have been easier to
fly, but because of their poor maneuverability they could not be brought
up into the air. By using their insight and skillful experiments the Wright
brothers made the first successful flight with Kitty Hawk in 1905. The fact
that this plane was unstable was a strong impetus for the development
of autopilots based on feedback.
Autopilots
Since it was quite tiresome to fly an unstable aircraft, there was strong
motivation to find a mechanism that would stabilize an aircraft. Such a
device, invented by Sperry, was based on the concept of feedback. Sperry
used a gyro-stabilized pendulum to provide an indication of the vertical.
He then arranged a feedback mechanism that would pull the stick to
make the plane go up if it was pointing down and vice versa. The Sperry
25
Chapter 1. Introduction
Figure 1.10 Picture from Sperry’s contest in Paris. Sperry’s son Lawrence is at
the stick and the mechanic walks on the wings to introduce disturbances. Notice
the proximity to the ground.
Autonomous Systems
Fully automatic flight including take off and landing is a development
that naturally follows autopilots. It is quite surprising that this was done
as early as 1947, see Figure 1.11. The flight was manually supervised but
the complete flight was done without manual interaction.
Autonomous flight is a challenging problem because it requires auto-
matic handling of a wide variety of tasks, landing, flying to the normal
flight altitude, navigation, approaching the airfield, and landing. This re-
quires a combination of continuous control with discrete control and logic,
so called hybrid control. In the flight by Robert E. Lee the logic required
was provided by an IBM computer with punch cards. The theory of hybrid
systems is still in its infancy. There are many similarities with the prob-
lem of designing humanoid robots. They require facilities for navigation,
map making, adaptation, learning, reasoning. There are many unsolved
26
1.7 Aeronautics
Figure 1.11 Excerpt from article in New York Times on September 23, 1947,
describing the first fully automatic transatlantic flight.
27
Chapter 1. Introduction
CM
CP
CP
CM
Figure 1.12 Schematic digram of two aircrafts. The aircraft above is stable be-
cause it has the center of pressure behind the center of mass. The aircraft below is
unstable.
28
1.8 Electronics and Communication
PSfrag replacements R2
R1
I
+
V
−
V1 V2
tive feedback amplifier. Before dealing with this we will, however, briefly
discuss an application of positive feedback.
V2 = GV (1.3)
Eliminating the variable V between Equations (1.2) and (1.3) gives the
following equation for the ratio of the output and input voltages
V2 R1
=G (1.4)
V1 R1 + R2 − GR2
This equation gives the gain of the amplifier with positive feedback. The
gain is very large if the resistors are chosen so that R1 + R2 − GR2 is
small. Assume for example that R1 = 100kΩ , R1 = 24kΩ , and G = 5.
The formula above shows that the gain of the feedback system is 25. With
R2 = 24.5kΩ the gain will be 50, and with R1 = 25kΩ the gain is infinite.
29
Chapter 1. Introduction
R2
PSfrag replacements
R1
I
−
V
+
V1 V2
Regeneration was the word used for feedback at the time, and Arm-
strongs amplifier was called a super-regenerative amplifier because it
obtained high gain by positive feedback. Armstrongs invention made it
possible to build inexpensive amplifiers with very high gain. The ampli-
fiers were, however, extremely sensitive and they could easily start to
oscillate. Prices of vacuum tubes also dropped and the interest in the
amplifier dwindled. Next we will discuss another use of feedback in an
amplifier which still has profound consequences.
Eliminating the variable V between Equations (1.2) and (1.5) gives the
30
1.8 Electronics and Communication
following equation for the ratio of the output and input voltages
V2 R2 1
=− R2
(1.6)
V1 R1 1 + 1
G (1 + R1 )
This equation gives the gain of the amplifier with negative feedback. Since
the gain G is a very large number, typically of the order of 105 or 108, it
follows from this equation that the input-output property of the amplifier
is essentially determined by resistors R1 and R2 . These are passive com-
ponents which are very stable. The properties of the active components
appear in parameter G. Even if G changes significantly, the input-output
gain remains constant. Also notice that the relation between V out and Vin
is very close to linear even if the relation between V out and V , equation
1.5, is strongly nonlinear.
Like many clever ideas the idea of the feedback amplifier seems almost
trivial when it is described. It took, however, six years of hard work for
Black to come up with it. The invention and development of the feedback
amplifier was a key step in the development of long distance communi-
cations. The following quote from the presentation of the IEEE Lamme
Medal to Black in 1957 gives a perspective on the importance of Black’s
invention of the feedback amplifier:
It is no exaggeration to say that without Black’s invention, the
present long-distance telephone and television networks which
cover our entire country and the transoceanic telephone cables
would not exist. The application of Black’s principle of negative
feedback has not been limited to telecommunications. Many of
the industrial and military amplifiers would not be possible
except for its use. ... Thus, the entire explosive extension of
the area of control, both electrical and mechanical, grew out of
an understanding of the feedback principle. The principle also
sheds light in psychology and physiology on the nature of the
mechanisms that control the operation of animals, including
humans, that is, on how the brain and senses operate.
It is interesting to observe that while Armstrong used positive feedback
Black was using negative feedback.
Feedback quickly became an indispensable companion of electronics
and communication. The applications are abundant. Today we find inter-
esting use of feedback in power control in system for cellular telephony.
A handset must naturally use enough power to transmit so that it can be
heard by the nearest station, using too much power will increase interfer-
ence with other handsets, necessitating use of even more power. Keeping
the power at the correct level gives a large pay-off because the batteries
will last longer.
31
Chapter 1. Introduction
Figure 1.15 Circuit diagram of William Hewletts oscillator that gives a stable
oscillation through nonlinear feedback using a lamp.
32
1.9 Automotive
is in operating range. The tracking and the sled servos are also used to
switch between tracks. The servos are all based on error feedback since
only the error signal is available from the sensors. The major disturbance
is due to misalignment of the track. In a CD player this is due to an
off set of the center both due to manufacturing variations and errors in
centering of the CD. The disturbance is approximately sinusoidal. The
tracking accuracy of the systems is quite remarkable. For a typical CD
player, which stores 650 M-bytes, the variations in a track are typically
200 µ m, the track width is 1.6 µ m and the tracking accuracy is 0.1µ m.
The tracking speed varies from 1.2 to 1.4m/s. The servos used in the
Digital Versatile Disc (DVD) and in optical memories are very similar to
the servos in a CD but the precision is higher. For a Digital Versatile Disc
(DVD) the variations in a track are typically 100 µ m, the track width is
1.6 µ m and the tracking accuracy is 0.022µ m. The tracking speed varies
in the range 3.5m/s. The quality of the major servo loops have a direct
impact on the performanc of the storage system. A better tracking servo
permits a higher storage density in an optical drive. An improved servo for
switching tracks is immediately reflected in the search time in an optical
memory.
1.9 Automotive
Cars are increasingly being provided with more and more control systems.
The possibility of doing this are due to availability of cheap microproces-
sors and sensors and good control technology. The automotive industry has
33
Chapter 1. Introduction
Reducing Emissions
California introduced a standard that required a substantial reduction of
emissions for internal combustion engines. To achieve this it was neces-
sary to introduce feedback in the engine based on measurement of the
oxygen in the exhaust. The following quote from a plenary lecture by
William E. Powers, (vice president at Ford) at the 1999 World Congress
of IFAC is illuminating.
The automobiles of the 1990s are at least 10 times cleaner and
twice as fuel efficient as the vehicles of the 1970s. These ad-
vancements were due in large part to distributed microproces-
sor based control systems. Furthermore the resultant vehicles
are safer, more comfortable and more maneuverable.
The advances were based on feedback control of the airfuel ratio.
34
1.9 Automotive
Technology Drivers
The automotive industry is an important driver for technology because of
the large number of produced parts and hard requirements for low cost.
Several interesting developments took place when computers started to be
used for computer control of engines. To save costs the microcomputer and
input-output devices that connect sensors and actuators were merged on
one chip, so called micro controllers. These devices made computer control
cost effective in many other fields. The total number of micro-controllers
for embedded systems now far exceed the number of microprocessors man-
ufactured each year. The automotive applications also required new sen-
sors and actuators. New accelerometers and gyros based on MEMS devices
35
Chapter 1. Introduction
Figure 1.19 The Mercedes A-class is a small car where control helped to solve a
serious problem.
Autonomous Driving
There have been several attempts at developing autonomous vehicles. In
1995 Dickmanns demonstrated a fully autonomous Mercedes Benz with
vision sensors. Under human supervision the car drove autonomously
on the Autobahn from Munich to Copenhagen. Experiments with au-
tonomous platoon driving have also been done in California in the Path
program. It has been demonstrated experimentally that it is possible to
drive a platoon of cars at high speed with only a meters separation bete-
ween the cars. Automatic attachement and separation of cars to the pla-
toon have also been demonstrated.
1.10 Computing
There has been a strong symbiosis between control and computing. Com-
puting devices are integral parts of a controller and computing and sim-
ulation are used extensively in design and validation of a control system.
Analog Computing
Early controllers, such as the centrifugal governor, were implemented
using mechanical devices. Integral action was implemented using the ball
and disc integrator invented by Lord Kelvin. In the process industries
36
1.10 Computing
analog computing was instead done using pneumatic devices. The key
elements were pneumatic amplifiers, restrictions and volumes. Feedback
was used extensively to obtain linear behavior from the nonlinear devices.
The early development of control was severely hampered by the lack of
computing, many clever graphical methods were developed to obtain in-
sight and understanding using modest computing. The situation is sum-
marized very clearly in the following quote from Vannevar Bush from
1923.
Bush later built the the first mechanical differential analyzer. The key
elements were the ball and disc integrator and the torque amplifier. It
could be used to integrate a handful of differential equations. This com-
puter was used by Ziegler and Nichols to devise tuning rules for PID
controllers.
Bush’s work laid the foundation for analog computing which developed
rapidly when cheap electronic amplifiers became available. This coincided
with the emergence of control and analog computing became the standard
tool for simulation of control systems. The analog computers were however
large expensive systems that required a large staff to maintain. The use
of analog computing was thus limited to persons with access to these rare
resources. Analog computing was also used to implement the controllers
in the servomechanism era and through the 1960s. Controllers were also
implemented as small dedicated analog computers.
Even if computer control is the dominating technology for implement-
ing controllers there are still niches where analog computing is used ex-
tensively. One area is micro-mechanical systems (MEMS) where mechan-
ics and control are integrated on the same chip. Analog computing is also
used for systems with extremely fast response time.
Computer Control
When digital computers became available they were first used for comput-
ing and simulation. The early computers were large and expensive and
not suitable to be embedded in controllers. The first computer controlled
systems was installed by TRW at the Port Arthur refinery in Texas in
1959. This initiated a development which started slowly and accelerated
37
Chapter 1. Introduction
Simulation
Simulation is an indispensable tool for the control engineer. Even if sys-
tems can be designed based on relatively simple models it is essential
to verify that the system works in a wide range of operations. This is
typically done by simulating the closed loop system. To be reliable the
simulation requires a high fidelity model of the system. Development of
such model is often time consuming. Sometimes parts of the real system
is actually interfaced with the simulator, so called hardware in the loop
simulation. Vehicle and flight simulators are typical examples. If a con-
troller is build using a dedicated computer it can also be verified against
the simulation. Simulation can also be used for many other purposes, to
explore different systems configurations, for operator training and for di-
agnostics. Because of the advances in computers and software simulation
is now easily available at every engineers desk top.
The Internet
The Internet was designed to be an extremely robust communication net-
work. It achieves robustness by being distributed and by using feedback.
The key function of the system is to transmit messages from a sender to
a receiver. The system has a large number of nodes connected with links.
At each node there are routers that receives messages and sends them
out to links. The routers have buffers that can store messages. It is desir-
able to operate the system to exploit capacity by maximizing throughput
subject to the constraint that all users are treated fairly. There are large
variations in traffic and in the lengths of the messages. Routers and links
38
1.11 Mathematics
can also fail so the system may also be changing. The Internet depends
critically on feedback to deal with uncertainty and variations. All control
is decentralized
The routers receive messages on the incoming lines and distributes
them to the outgoing lines. Messages are stored in a buffer if the router
can not handle the incoming traffic. In the simplest scheme a router that
has a full buffer will simply drop the incoming messages. Information
about congestion is propagated to the sender indirectly through the lost
messages.
Flow control is done by the senders. When a message is received by a
receiver it sends an acknowledgment to the sender. The sender can detect
lost messages because the messages are tagged. A very simple algorithm
for traffic control was proposed by Jacobson 1990. This algorithm is called
Additive Increase Multiplicative Decrease (AIMD) works as follows. The
transmission rate is increased by additively as long as no messages are
lost. When a message is lost the transmission rate is reduced by a factor
of 2.
The Internet is a nice illustration that a very large distributed system
can be controlled effectively by reasonably simple control schemes. There
are many variations of the basic scheme described above. Many technical
details have also been omitted. One drawback with the current scheme
is that the control mechanism creates large variations in traffic which is
undesirable. There are many proposals for modifications of the system.
1.11 Mathematics
There has always been a strong interplay between mathematics and con-
trol. This has undoubtedly contributed to the success of control. The the-
ory of governors were developed by James Clarke Maxwell in a paper
from 1868, about 100 years after James Watt had developed the governor.
Maxwell realized that stability was connected to the algebraic problem
of determining if an algebraic equation has roots in the right half plane.
Maxwell turned to the mathematician Routh to get help. An analogous
situation occurred in the development of water turbines where Stodola
turned to the mathematician Hurwitz for assistance.
Mathematics played a major role in the development of servomecha-
nism theory in the 1940s. There were a number of outstanding mathemati-
cians at the Radiation Laboratory at MIT, there were also outstanding
mathematicians at Bell Laboratories at the time. An interesting perspec-
tive can be obtained by comparing the major advances in the theory of
feedback amplifiers with the meager advances in process control and to
speculate what could have happened if there had been mathematicians
39
Chapter 1. Introduction
Numerical Mathematics
A standard computational problem is to solve ordinary differential equa-
tions by time-stepping methods. As solutions often vary significantly over
the range of integration, efficient computation requires step length con-
trol. This is done by estimating the local error and adjusting the step
length to make this error sufficiently small. Most solvers for differential
equations have carried out the correction in a simplistic fashion, and the
adjustment has often been mixed with other algorithmic elements.
Recently a drastic improvement has been made by viewing step length
adjustment as a feedback problem. Substantial improvements of perfor-
mance can be obtained by replacing the heuristic schemes for step length
adjustment, that were used traditionally, with a scheme based on a PID
controller. These advantages are achieved without incurring additional
computational costs. This has resulted in more reliable software, as well as
software with much better structure. Knowledge of basic control schemes
has thus proven very beneficial. It is likely that the same idea can be
applied to other numerical problems.
40
1.12 Physics
1.12 Physics
Adaptive Optics
A severe problem in astronomy is that turbulence in the atmosphere blurs
images in telescopes because of variations in diffraction of light in the at-
mosphere. The blur is of the order of an arc-second in a good telescope.
One way to eliminate the blur is to move the telescope outside the Earths
atmosphere as is done with the Hubble telescope. Another way is to use
feedback to eliminate the effects of the variations in a telescope on the
Earth. This is the idea of adaptive optics. A schematic picture of a sys-
tem for adaptive optics is shown in Figure 1.20. The reference signal is
a bright star or an artificial laser beam projected into the atmosphere.
The actuator which is shown simply as a box in the figure is obtained
by reflecting the light on a mirror that can be deformed selectively. The
mirror can have from 13 to 1000 elements. The error signal is formed by
analyzing the shape of the distorted wave form from the reference. This
signal is sent to the controller which adjusts the deformable mirror. The
light from the observed star is compensated because it is also reflected in
the deformable mirror before it is sent to the detector. The wave lengths
used for observation and control are often different. Since diffraction in
the atmosphere changes quite rapidly the response time of the control
41
Chapter 1. Introduction
Quantum Systems
Control of quantum systems is currently receiving a lot of interest. Molec-
ular dynamics is a very spectacular application. The idea is to use modu-
lated laser light to break up bonds in molecules to obtain ions which can
react with other ions to form new molecules. This is done by tailoring the
laser pulses so that they will break specific bonds between atoms. This is
precision surgery at the molecular level, quite different from the methods
used in conventional chemistry.
1.13 Biology
42
1.13 Biology
Human Posture
The human body has many feedback systems. They allow us to stand up-
right, to walk, jump and balance on ropes. They also adjust the sensitivity
of our eyes and ears, enabling us to see and hear over a wide range of
intensity levels. They maintain a constant body temperature and a del-
icate balance of chemical substances in our body. As an illustration we
will discuss the system that allows us to stand upright. The key features
of the system are known although several details are poorly understood.
The primary sensors are the semicircular canals located in the mastoid
bone close to the ear. The sensors consist of toroidal canals that are filled
with liquid. Neurons connected to hairs in the canals give signals related
to the motion of the head. The major actuators are muscles in feet, legs,
knees, hips and arms. There are also sensory neurons in the feet and the
muscles. There is local feedback from pressure sensors in the feet and
sensors in the muscles to the actuating muscles. This loop has a reaction
time of about 20 ms. The interconnection between sensors and actuators
is made in the spinal cord. These interconnections are responsible for fast
feedback like reflexes. The reaction time is of the order of 100 ms. There is
also a high level feedback loop that receives information from the vestibu-
lar system, which gives information about the position and orientation of
our body parts in space. The sensory information is processed in the cere-
bellum and transmitted to the muscle neurons in the spinal cord. This
feedback has a reaction time of about 250 ms.
This system for control of posture illustrates that feedback can be used
to stabilize an unstable system. It also shows that there are very reliable
biological feedback systems which are essential to everyday life. A par-
ticularly interesting feature is that the system has learning capabilities.
Think about a child learning to stand up and walk or learning to bicycle.
These functions are far superior to those of any technical system.
A Simple Experiment
A simple experiment on one of the systems in the body can be executed
manually with very modest equipment. Take a book with text and hold it
in front of you. Move the text sideways back and forth and increase the
speed of motion until the text is blurred. Next hold the text in front of you
and move the head instead. Notice the difference in the speeds when the
text gets blurred. You will observe that higher speeds are possible when
you move your head. The reason for this is that when you move the text,
the information about the motion comes via the processing of the image
43
Chapter 1. Introduction
at your retina, but when you move your head the information comes from
the semicircular canals. The feedback from the visual processing is much
slower because it uses higher functions in the brain.
There are many other nice control systems in the human body. Top
performing athletes such as tennis players have interesting abilities for
very advanced motion control that involves much interaction with vision.
Humans also have interesting learning abilities.
1.14 Summary
This chapter has given glimpses of how control and feedback have influ-
enced the development of technology and how control is used. The ex-
amples show that control has emerged concurrently with new technology,
that it has had a major influence on technology and that it sometimes has
been an enabler. A large number of control systems ranging from small
micro devices to large global systems for generation and distribution of
electricity and large communication systems have also been given. The
wide range of uses of control also point to some difficulties. The ideas of
control and feedback are abstract which makes them less obvious. It is
therefore common that the ideas are neglected in favor of hardware which
is much easier to talk about.
44
2
Feedback
2.1 Introduction
Feedback is a powerful idea, which is used extensively in natural and
technical systems. The principle of feedback is very simple: base correct-
ing actions on the difference between desired and actual performance.
In engineering feedback has been rediscovered and patented many times
in many different contexts. Use of feedback has often resulted in vast
improvements in system capability, sometimes they have even be revolu-
tionary as discussed in Chapter 1. The reason for this is that feedback
has some truly remarkable properties. In this chapter we will discuss
some of the properties of feedback that can be understood intuitively. The
benefits of feedback can often be obtained using simple forms of feedback
such as on-off control and PID control, which are discussed in Section 2.2.
Particular attention is given to integral action which is has truly remark-
able properties. Feedback systems may appear complicated because they
involve many different subsystems and many different technologies. To
reduce the complexity it is necessary to have abstractions that makes it
possible to have an overview of the systems. Section 2.3 presents differ-
ent ways to describe feedback systems. The block diagram is the most
important representation because it is a uniform description that can be
adapted to many different purposes. The remarkable properties of feed-
back are presented in Section 2.4. It is shown that feedback can reduce the
effects of disturbances and process variations, it can create well defined
relations between variables, it makes it possible to modify the properties
of a system, e.g. stabilize a unstable system. The discussion is based on
block diagrams and simple static mathematical models. The major draw-
back is that feedback can create instability. This is discussed briefly in
Section 2.5. To understand stability it is necessary to have knowledge of
45
Chapter 2. Feedback
A u B u C u
e e e
Figure 2.1 Controller characteristics for ideal on-off control (A), and modifications
with dead zone (B) and hysteresis (C).
On-Off Control
The feedback can be arranged in many different ways. A simple feedback
mechanism can be described as follows:
(
umax , if e > 0
u= (2.1)
umin , if e < 0
46
2.2 Simple Forms of Feedback
PID Control
The reason why on-off control often gives rise to oscillations is that the
system over reacts since a small change in the error will make the ma-
nipulated variable change over the full range. This effect is avoided in
proportional control where the characteristic of the controller is propor-
tional to the control error for small errors. This can be achieved by making
the control signal proportional to the error
u = k(r − y) = ke (2.2)
Integral Action Proportional control has has the drawback that the
process variable often deviates from its reference value. This can be avoided
by making the control action proportional to the integral of the error
Zt
u(t) = ki e(τ )dτ (2.3)
0
where ki is the integral gain. This control form is called integral control. It
follows from this equation that if there is a steady state where the control
signal and the error are constant, i.e. u(t) = u0 and e(t) = e0 respectively
then
u0 = ki e0 t
This equation is a contradiction unless e0 = 0. It has thus demonstrated
that there will be no steady state error with a controller that has integral
action. Notice that the argument also holds for any process and any con-
troller that has integral action. The catch is that there may not always be
a steady state because the system may be oscillating. This amazing prop-
erty which we call the Magic of Integral Control has been rediscovered
many times. It is one of the properties that have strongly contributed to
the wide applicability of PID control.
de(t)
e(t + Td ) e(t) + Td ,
dt
which predicts the error Td time units ahead, see Figure 2.2. Combining
47
Chapter 2. Feedback
Figure 2.2 A PID controller takes control action based on past, present and future
control errors.
Zt
de(t)
u(t) = ke(t) + ki e(τ )dτ + kd
dt
0
(2.4)
Zt
1 de(t)
= k e(t) + e(τ )dτ + Td
Ti dt
0
The control action is thus a sum of three terms representing the past by
the integral of the error (the I-term ), the present (the P-term ) and the
future by a linear extrapolation of the error ( the D-term ). The term
e + Td de/dt is a linear prediction of the error Td time units in the future.
Notice that the controller can be parameterized in different ways. The
second parameterization is commonly used in industry. The parameters
of the controller are called: are proportional gain k, integral time Ti , and
derivative time Td .
The PID controller is very useful. It is capabable of solving a wide
range of control problems. The PI controller is the most common con-
troller. It is quoted that about 90% of all control problems can be solved by
PID control, many of these controllers are actually PI controller because
derivative action is not so common. There are more advanced controllers
which differ from the PID controller by using more sophisticated methods
for prediction.
48
2.3 Representation of Feedback Systems
Schematic Diagrams
In all branches of engineering, it is common practice to use some graphical
description of systems. They can range from stylistic pictures to drasti-
cally simplified standard symbols. These pictures make it possible to get
an overall view of the system and to identify the physical components.
Examples of such diagrams are shown in Figure 2.3
Block Diagrams
The schematic diagrams are useful because they give an overall picture of
a system. They show the different physical processes and their intercon-
nection, and they indicate variables that can be manipulated and signals
that can be measured.
A special graphical representation called block diagrams has been de-
veloped in control engineering. The purpose of block diagrams is to empha-
size the information flow and to hide technological details of the system.
It is natural to look for such representations in control because of its mul-
tidisciplinary nature. In a block diagram, different process elements are
shown as boxes. Each box has inputs denoted by lines with arrows point-
ing toward the box and outputs denoted by lines with arrows going out of
the box. The inputs denote the variables that influence a process and the
outputs denote some consequences of the inputs that are relevant to the
feedback system.
Figure 2.4 illustrates how the principle of information hiding is used
to derive an abstract representation of a system. The upper part of the
picture shows a photo of a physical system which a small desk-top process
in a control laboratory. It consists of two tanks, a pump that pumps wa-
ter to the tanks, sensors, and a computer which implements the control
algorithm and provides the user interface. The purpose of the system is
to maintain a specified level in the lower tank. To do so, it is necessary
to measure the level. The level can be influenced by changing the speed
of the motor that pumps water into the upper tank. The voltage to the
49
Chapter 2. Feedback
amplifier that drives the pump is selected as the control variable. The
controller receives information about the desired level in the tank and
the actual tank level. This is accomplished using an AD converter to con-
vert the analog signal to a number in the computer. The control algorithm
in the computer then computes a numerical value of the control variable.
This is converted to a voltage using a DA converter. The DA converter is
connected to an amplifier for the motor that drives the pump.
The first step in making a block diagram is to identify the impor-
tant signals: the control variable, the measured signals, disturbances and
goals. Information hiding is illustrated in the figure by covering systems
by a cloth as shown in the lower part of Figure 2.4. The block diagram is
simply a stylized picture of the systems hidden by the cloth.
In Figure 2.4, we have chosen to represent the system by two blocks
only. This granularity is often sufficient. It is easy to show more details
simply by introducing more subsystems, as indicated in Figure 2.5 where
we show the drive amplifier, motor, pump, and tanks, the sensors with
electronics, the AD converter, the computer and the DA converter. The
50
2.3 Representation of Feedback Systems
Figure 2.4 Illustrates the process of information hiding used to obtain a block
diagram. The top figure is a picture of the physical system, the middle figure is
obtained by hiding many details about the system and the bottom figure is the
PSfrag replacements
block diagram.
Figure 2.5 A more detailed block diagram of the system in Figure 2.4 showing
controller C, amplifier A, pump, tanks and sensor.
51
Chapter 2. Feedback
A B
C D
PSfrag replacements
Figure 2.6 A simple hydraulic system with an inflow and a free outflow is shown
in A. The block diagram representation of the system is shown in B. The system
obtained by connecting two hydraulic systems is shown in C. This system cannot be
represented by the series connection of the block diagrams in B.
Causality
The arrows in a block diagram indicate causality because the output of
a block is caused by the input. To use the block diagram representation,
it is therefore necessary that a system can be partitioned into subsys-
tems with causal dependence. Great care must be exercised when using
block diagrams for detailed physical modeling as is illustrated in Fig-
ure 2.6. The tank system in Figure 2.6B is a cascade combination of the
two tanks shown in Figure 2.6B. It cannot be represented by cascading
the block diagram representations because the level in the second tank
influences the flow between the tanks and thus also the level in the first
tank. When using block diagrams it is therefore necessary to choose blocks
to represent units which can be represented by causal interactions. We
can thus conclude that even if block diagrams are useful for control they
also have serious limitation. In particular they are not useful for serious
physical modeling which has to be dealt with by other tools which permit
bidirectional connections.
Examples
An important consequence of using block diagrams is that they clearly
show that control systems from widely different domains have common
features because their block diagrams are identical. This observation was
one of the key factors that contributed to the emergence of the discipline
of automatic control in the 1940s. We will illustrate this by showing the
block diagrams of some of the systems discussed in Chapter 1.
52
2.3 Representation of Feedback Systems
Centrifugal governor
−1
53
Chapter 2. Feedback
Autopilot
−1
PSfrag replacements
R2
R1
I
−
V
+
V1 V2
Even if block diagrams are simple, it is not always entirely trivial to obtain
them. It happens frequently that individual physical components to not
necessarily correspond to specific blocks and that it may be necessary to
use mathematics to obtain the block. We illustrate this by an example.
where G is the gain of the amplifier and the negative sign indicates neg-
ative feedback. If the current I into the amplifier is negligible the current
through resistors R1 and R2 are the same and we get
V1 − V V − V2
=
R1 R2
54
PSfrag replacements
PSfrag replacements
R1
2.3 Representation of Feedback Systems
PSfrag Rreplacements
1 + R2
−G
V1 e V V2
R1
V1 −G Σ R1 + R 2 −G
e
V R1
V2 R2
R1
R2
V2 V1 e V − V2
G R1 R1
− V2 R2 Σ R1 + R 2 G
− RR12
− RR1
2
V1 V − V2
R1 R1
R2
Σ R1 + R 2 G
−1
Figure 2.10 Three block diagrams of the feedback amplifier in Figure 2.9.
Solving this equation for the input voltage V to the amplifier we get
R2 V 1 + R 1 V 2 R2 R1
V = = V1 + V2
R1 + R 2 R1 + R 2 R2
This equation can be represented by one block with gain R2 /( R1 + R2 )
and the input V1 + R1 V2/ R1 and we obtain the block diagram shown in
Figure 2.10. The lower representation where the process has positive gain
and the feedback gain is negative has become the standard of represent-
ing feedback systems. Notice that the individual resistors do not appear
as individual blocks, they actually appear in various combinations in dif-
ferent blocks. This is one of the difficulties in drawing block diagrams.
Also notice that the diagrams can be drawn in many different ways. The
middle diagram in Figure 2.10 is obtained by viewing − V2 as the output
of the amplifier. This is the standard convention where the process gain
is positive and the feedback gain is negative. The lowest diagram in Fig-
ure 2.10 is yet another version, where the ratio R1 / R2 is brought outside
the loop. In all three diagrams the gain around the loop is R2 G /( R1 + R2 ),
this is one of the invariants of a feedback system.
55
PSfrag replacements
Chapter 2. Feedback
d n
r e u v x y
Σ C Σ P Σ
−1
Controller Process
tems, their block diagrams are identical apart from the labeling of blocks
and signals, compare Figures 2.7, 2.8 and 2.10. This illustrates the uni-
versality of control. A generic representation of the systems is shown in
Figure 2.11. The system has two blocks. One block P represents the pro-
cess and the other C represents the controller. Notice negative sign of the
feedback. The signal r is the reference signal which represents the desired
behavior of the process variable x.
Disturbances are an important aspect of control systems. In fact if
there were no disturbances there is no reason to use feedback. In Fig-
ure 2.11 there are two types of disturbances, labeled d and n. The distur-
bance labeled d is called a load disturbance and the disturbance labeled
n is called measurement noise. Load disturbances drive the system away
from its desired behavior. In Figure 2.11 it is assumed that there is only
one disturbance that enters at the system input. This is called an input
disturbance. In practice there may be many different disturbances that
enter the system in many different ways. Measurement noise corrupts the
information about the process variable obtained from the measurements.
In Figure 2.11 it is assumed that the measured signal y is the sum of the
process variable x and measurement noise. In practice the measurement
noise may appear in many other ways.
The system in Figure 2.11 is said to have error feedback, because the
control actions are based on the error which is the difference between
the reference r and the output y. In some cases like a CD player there
is no explicit information about the reference signal because the only
information available is the error signal. In such case the system shown
in Figure 2.11 is the only possibility but if the reference signal is available
there are other alternatives that may give better performance.
56
2.3 Representation of Feedback Systems
d n
r s e u v x y
F Σ C Σ P Σ
−1
Controller Process
Figure 2.12 Block diagram of a generic feedback system with two degrees of
freedom.
57
Chapter 2. Feedback
PSfrag replacements
r e R y
Σ
− y2
Mathematical Models
A block diagram gives an overview of a system but more details are re-
quired to obtain a more complete description of a system. In particular it is
necessary to describe the behavior of each individual block. This requires
mathematical models. A function
y = f (u)
58
2.4 Properties of Feedback
10
5
y
0
PSfrag replacements
−5
0 0.5 1 1.5 2 2.5 3 3.5 4
u
Figure 2.14 Static input-output function for a system and the linear approxima-
tion around the operating point u0 = 3.5. The slope of the curve at the operating
point is the gain of the system.
of the tangent is called the gain of the system at the operating point.
Assume that the control variable at the operating point has the value u0 .
The corresponding value of the output is then y0 = f (u0). For values of the
control variable close to the operating point the system can approximately
be described by the model
y = Ku
which we call a linear static model. Static models of control systems have
severe limitations, because many of the properties of feedback systems
rely on dynamic effects. A major part of this book will be devoted to this
beginning with next chapter which deals with dynamics. Control is thus
intimately connected with dynamics.
Static Analysis
We will start by a very simplistic analysis of a system with error feedback.
Consider the system in Figure 2.11. Assume that the variables r, d and n
are constants and that the process and the controller can be described by
linear static models. Let kp be the process gain and kc the controller gain.
59
Chapter 2. Feedback
The following equations are obtained for the process and the controller.
y= x+n
x = kp(u + d) (2.5)
u = kc (r − y)
The product L = kp kc is called the loop gain. It is the total gain around the
feedback loop. It is an important system property which is dimension-free.
Several interesting conclusions can be drawn from (2.6). First we ob-
serve that since the equation is linear we can discuss the effects of refer-
ence values r, load disturbances d and measurement noise n separately.
It follows from (2.6) that the output will be very close to the reference
value if the loop gain L = kp kc is large. It also follows from (2.6) that the
effect of the load disturbances will be small if the controller gain is large.
We will now take a closer look at some of the responses.
Assuming that r = 0 and n = 0 we find that the process variable is
given by
kp
x= d
1 + kp kc
The influence of the load disturbance on the output can be reduced sig-
nificantly by having a controller with high gain.
If the disturbances are zero, i.e. d = n = 0 the response to reference
signals is given by
kp kc
x= r (2.7)
1 + kp kc
By having a controller with high gain the process variable will be very
close to the reference value. For example if kp kc = 100 the deviation
between x and r is less than 1%
The properties of the process are seldom constant. To investigate the
effects of process variations we differentiate (2.7) with respect to the pro-
cess gain kp. This gives
dx kc kc 1 x 1
= r= r=
dkp (1 + kp kc )2 1 + kp kc 1 + kp kc kp 1 + kp kc
60
PSfrag replacements
r e u y
Σ C P
1
−
k
Figure 2.15 This feedback system has the input output relation y = kr even if
the process P is highly nonlinear.
Hence
d log x dx/ x 1
= = (2.8)
d log kp dkp/ kp 1 + kp kc
The relative variation in the process variable caused by process variations
will thus be very small if the loop gain is high. For example if the loop
gain is kp kc = 100 it follows from (2.8) that a 10% variation in the process
gives only a variation of 0.1% in the relation between the process variable
and the reference. This was the idea used by Black when he invented the
feedback amplifier, (2.8) was actually part of Black’s patent application.
The simple analysis above captures several properties of feedback. The
analysis can however be misleading because it does not consider the dy-
namics of the process and the controller. The most serious drawback is
that most systems will become unstable when the loop gain is large. An-
other factor is that the trade-off between reduction of load disturbances
and injection of measurement noise is not represented well. In practice
the load disturbances are often dominated by components having low fre-
quencies while measurement noise often has high frequencies.
y = f (u)
u = k(r − y)
Eliminating u between these equations we find that the closed loop system
is then described by
1
y + f −1( y) = r
k
61
Chapter 2. Feedback
0.8
0.6
y
0.4
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
u, r
PSfrag replacements
Figure 2.16 Input output relations for the process y = u5 (solid lines) and for the
feedback system in Figure 2.15 (dashed lines) when the controller gain k is 100.
xr vr Ir I v x
Cp Cv CI
Controller Motor
62
2.4 Properties of Feedback
63
Chapter 2. Feedback
EXAMPLE 2.12—HAPTICS
There are special types of joy sticks which are provided with motors so
that the joy stick can generate forces that give additional information to
the user. With these joy sticks it is possible to use feedback to generate
forces when the cursor approaches given regions. In this way it is possible
to simulate things like pushing on soft objects.
The possibility of using feedback to modify the behavior of systems is an
idea that has very wide applicability.
2.5 Stability
Static analysis of feedback systems is based on the assumption that a con-
trol action is immediately manifested by a change in system output. This
is a strong simplification because systems are typically dynamic which
means that changes in the input do not immediately give rise to changes
in the output. A major effort of control theory is actually devoted to finding
appropriate ways to describe dynamical phenomena.
One consequence of the systems being dynamic is that feedback may
give rise to oscillations. The risk for instability is the major drawback of
using feedback. The static analysis resulting in (2.6) has indicated that
it is advantageous to have controllers with high gain. It is a common
practical experience that feedback systems will oscillate if the feedback
gain is too high. After a perturbation the control error typically has one
of the following behaviors:
64
2.5 Stability
Stable Unstable
1 1
0.5 0.5
0 0
0 10 20 30 40 50 0 10 20 30 40 50
Stable Unstable
1 1
0 0
PSfrag replacements
−1 −1
0 10 20 30 40 50 0 10 20 30 40 50
This is illustrated in Figure 2.18, the behaviors S1 and S2 are called stable
the ones labeled U1 and U2 are unstable. There are also intermediate
situations where the error remains constant or oscillates with constant
amplitude.
In a paper from 1868 Maxwell made the important observation that
stability was related to properties of the roots of an algebraic equation. In
particular he established the following relations between the behaviors in
Figure 2.18 and the properties of the roots.
65
Chapter 2. Feedback
−
−d
vr u v
PSfrag replacements
Contr Engine Σ Car
−d
vr u v
Σ Contr Engine Σ Car
−
Figure 2.19 Open and closed loop systems for cruise control. The disturbance is
the slope of the road.
66
2.6 Open and Closed Loop Systems
u = kcvr
e = vr − v = (1 − kp kc )vr + kp d (2.10)
kp kc 1
e = vr − v = vr + d (2.11)
1 + kp kc 1 + kp kc
67
PSfrag replacements
Chapter 2. Feedback
d
−F
u y
P1 Σ P2
r u y
Σ C Σ P1 Σ P2
−1
Figure 2.21 Control system that combines feedback and feedforward from a mea-
surable disturbance.
2.7 Feedforward
Despite all the nice properties of feedback, it also has some limitations.
When using feedback, an error must occur before the control system will
act. This means that there must be an error before corrective actions are
taken. Feedback is thus reactive. In some circumstances it is possible to
measure a disturbance before it enters the system. It then appears natu-
ral to use the information about the disturbance to take corrective action
before the disturbance has influenced the system. This way of controlling
a system is called feedforward. The concept of feedforward is illustrated
in Figure 2.20. The effect of the measured signal is reduced by measuring
it and generating a control signal that counteracts it. Since feedforward
attempts to match two signals it requires good process models, otherwise
the corrections may have the wrong size or it may be badly timed. Feed-
forward is often combined with feedback as is illustrated in Figure 2.21.
Such a system combines the good features of feedback and feedforward.
The system with a controller having two degrees of freedom shown in
Figure 2.12 can also be interpreted as a system that used feedforward to
68
2.8 Summary
Feedback Feedforward
Closed loop Open loop
Market driven Planning
Reactive Pro-active
Robust to modeling errors Sensitive to modeling errors
Risk for instability No risk for instability
2.8 Summary
The idea of feedback has been discussed in this section. It has been shown
that feedback is a powerful concept that has played a major role in the
development of many areas of engineering, e.g. process control, telecom-
munications, instrumentation and computer engineering. This widespread
use of feedback derives from its interesting properties.
• Feedback can reduce effects of disturbances
69
Chapter 2. Feedback
70
3
Dynamics
3.1 Introduction
From the perspective of control a dynamical system is such that the ef-
fects of actions do not occur immediately. Typical examples are: The ve-
locity of a car does not change immediately when the gas pedal is pushed.
The temperature in a room does not rise immediately when an air condi-
tioner is switched on. Dynamical systems are also common in daily life. An
headache does not vanish immediately when an aspirin is taken. Knowl-
edge of school children do not improve immediately after an increase of a
school budget. Training in sports does not immediately improve results.
Increased funding for a development project does not increase revenues
in the short term.
Dynamics is a key element of control because both processes and con-
trollers are dynamical systems. Concepts, ideas and theories of dynamics
are part of the foundation of control theory. Dynamics is also a topic of its
own that is closely tied to the development of natural science and math-
ematics. There has been an amazing development due to contributions
from intellectual giants like Newton, Euler, Lagrange and Poincare.
Dynamics is a very rich field that is partly highly technical. In this
section we have collected a number of results that are relevant for under-
standing the basic ideas of control. The chapter is organized in separate
sections which can be read independently. For a first time reader we rec-
ommend to read this section section-wise as they are needed for the other
chapters of the book. To make this possible there is a bit of overlap be-
tween the different sections. in connection with the other chapters. There
is a bit of overlap so that the different sections can be read independently.
Section 3.2 gives an overview of dynamics and how it is used in con-
trol which has inherited ideas both from mechanics and from electrical
71
Chapter 3. Dynamics
72
3.2 Two Views on Dynamics
box models. The external view is associated with names such as external
descriptions, input-output models or black box models. In this book we
will mostly use the words state models and input-output models.
dx1
= x1 − x13 − x2
dt
dx2
= x1
dt
which is a model of an electronic oscillator. The model (3.1) gives the ve-
locity of the state vector for each value of the state. These are represented
by the arrows in the figure. The figure gives a strong intuitive represen-
tation of the equation as a vector field or a flow. Systems of second order
can be represented in this way. It is unfortunately difficult to visualize
equations of higher order in this way.
The ideas of dynamics and state have had a profound influence on
philosophy where it inspired the idea of predestination. If the state of a
73
Chapter 3. Dynamics
3 M=1
x’=Mx−y−x
y’=x
1.5
0.5
0
y
−0.5
−1
−1.5
−2
Figure 3.1 Illustration of a state model. A state model gives the rate of change of
the state as a function of the state. The velocity of the state are denoted by arrows.
natural system is known ant some time its future development is complete
determined. The vital development of dynamics has continued in the 20th
century. One of the interesting outcomes is chaos theory. It was discovered
that there are simple dynamical systems that are extremely sensitive
to initial conditions, small perturbations may lead to drastic changes in
the behavior of the system. The behavior of the system could also be
extremely complicated. The emergence of chaos also resolved the problem
of determinism, even if the solution is uniquely determined by the initial
conditions it is in practice impossible to make predictions because of the
sensitivity of initial conditions.
The differential equation (3.1) is called an autonomous system be-
cause there are no external influences. Such a model is natural to use for
celestial mechanics, because it is difficult to influence the motion of the
planets. The situation in control is quite different because the external
influences are quite important. One way to capture this is to replace (3.1)
by
dx
= f ( x, u) (3.2)
dt
where u represents the effect of external influences. The model (3.2) is
called a controlled differential equation. The model implies that the veloc-
ity of the state can be influenced by the input u. Adding the input makes
the model richer. New questions arises, for example, what influence can
the control variable have on the trajectories of the system? Is it possible
to reach all points in the state space by proper choices of the control?
74
3.2 Two Views on Dynamics
Input Output
System
The function h is called the step response of the system. Notice that the
impulse response is the derivative of the step response.
Another possibility to describe a linear, time-invariant system is to
represent a system by its response to sinusoidal signals, this is called fre-
quency response. A rich powerful theory with many concepts and strong,
useful results have emerged. The results are based on the theory of com-
plex variables and Laplace transforms. The input-output view lends it
75
Chapter 3. Dynamics
dx
= f ( x, u)
dt (3.5)
y = g( x, u)
76
3.3 Linear Differential Equations
Standard Models
Standard models are very useful for structuring our knowledge. It also
simplifies problem solving. Learn the standard models, transform the
problem to a standard form and you are on familiar grounds. We will
discuss four standard forms
• Linear differential equations
• Transfer functions
• Frequency responses
• State equations
The first three standard forms are primarily used for linear time-invariant
systems. The state equations also apply to nonlinear systems.
dk d yk d yk
k
(α y1 + β y2) = α k1 + β k1
dt dt dt
If (u, y) is a pair of inputs and outputs it follows that (uP , yP) is also an
input output pair. Simlarly, if the (u1, y1) and (u2, y2) are pairs of inputs
and outputs that satisfy the Equation (3.6) α u1 + β u2, α y1 + β y2) is also
an input output pair, which is the principle of superposition.
dn y dn−1 y dn−2 y
+ a 1 + a 2 + . . . + an y = 0, (3.7)
dtn dtn−1 dtn−2
77
Chapter 3. Dynamics
α <0 aneg
3
1
2.5
0.8
2
0.6
y
y
1.5
PSfrag replacements 0.4 1
0.2 0.5
0 0
0 1 2 3 4 0 0.2 0.4 0.6 0.8 1
α >0 αt αt
n
X
y(t) = Ck eα k t, (3.9)
k=1
where Ck are arbitrary constants. The Equation (3.7) thus has n free
parameters.
eσ t sin ω t, eσ t cos ω t
which have oscillatory behavior, see Figure 3.4. The distance between zero
crossings is π /ω and corresponding amplitude change is eσπ /ω .
78
3.3 Linear Differential Equations
σ = −0.25ω σ = 0.25ω
40
0.6
30
0.4 20
0.2
y
y
10
PSfrag replacements 0 0
−0.2 −10
−0.4 −20
0 5 10 15 0 5 10 15
ωt ωt
Multiple Roots
When there are multiple roots the solution to Equation (3.7) has the form
n
X
y(t) = Ck (t) eα kt, (3.10)
k=1
Where Ck(t) is a polynomial with degree less than the multiplicity of the
root α k. The solution (3.10) thus has n free parameters.
dn y dn−1 y dn−2 y
n
+ a1 n−1 + a2 n−2 + . . . + an y = u(t) (3.11)
dt dt dt
has the solution
n
X Z t
y(t) = Ck−1(t) eα kt + h(t − τ )u(τ )dτ , (3.12)
k=1 0
dn h dn−1 h
+ a 1 + . . . + an h = 0 (3.13)
dtn dtn−1
with initial conditions
79
Chapter 3. Dynamics
The solution (3.12) is thus a sum of two terms, the general solution to
the homogeneous equation and a particular solution which depends on
the input u. The solution has n free parameters which can be determined
from initial conditions.
To show that (3.12) satisfies (3.11) we first observe that the sum in
(3.12) satisfies the homogeneous equation (3.7). Consider
Z t
v(t) = h(t − τ )u(τ )dτ ,
0
It follows from (3.13) and (3.14) that v satisfies the differential equation
(3.11).
dn y dn−1 y dn−2 y du
n
+ a1 n−1 + a2 n−2 + . . . + an y =
dt dt dt dt
Repeating this argument for higher derivatives we find that the Equa-
tion (3.6) has the solution
n
X Z t
α kt
y(t) = Ck−1(t) e + g(t − τ )u(τ )dτ , (3.15)
k=1 0
80
3.3 Linear Differential Equations
The solution is thus the sum of two terms, the general solution to the ho-
mogeneous equation and a particular solution. The general solution to the
homogeneous equation does not depend on the input and the particular
solution which depends on the input. The particular solution is given by
Z t
y(t) = g(t − τ )u(τ )dτ
0
n
X
g(t) = ck(t) eα kt. (3.17)
k=1
It thus has the same form as the general solution to the homogeneous
equation (3.10). The coefficients c k are given by the conditions (3.14). If
the characteristic equation has distinct roots c k(t) are constants. If α k is
a root of multiplicity m then c k(t) is a polynomial of degree m − 1.
The impulse response is also called the weighting function because the
second term of (3.15) can be interpreted as a weighted sum of past inputs.
The function H is called the unit step response or the step response for
short. It follows from the above equation that
dh(t)
g(t) = (3.20)
dt
81
Chapter 3. Dynamics
10
u(τ ) y (τ )
5
−5
0 1 2 3 4 5 6 7 8 9 10
1
g(10 − τ )
0.5
0
0 1 2 3 4 5 6 7 8 9 10
2
u(τ ) g(10 − τ )
PSfrag replacements 1
−1
τ 0 1 2 3 4 5 6 7 8 9 10
Figure 3.5 Illustration of the convolution integral for the impulse response g( t) =
e−4t . The top shows the input u in full lines and the output y in dashed lines. The
lower graphs illustrates how y (10) is obtained.
Z t
y(t) = g(t − τ )u(τ )dτ .
0
82
3.3 Linear Differential Equations
• The polynomial a(s) = s2 + a1 s + a2 has all its zeros in the left half
plane if all coefficients are positive.
83
Chapter 3. Dynamics
The number G (0) is called the static gain of the system because it tells
the ratio of the output and the input under steady state condition. If the
input is constant u = u0 and the system is stable then the output will
reach the steady state value y0 = G (0)u0. The transfer function can thus
be viewed as a generalization of the concept of gain.
Notice the symmetry between y and u. The inverse system is obtained
by reversing the roles of input and output. The transfer function of the
b(s) a(s)
system is and the inverse system has the transfer function .
a(s) b(s)
The roots of a(s) are called poles of the system. The roots of b(s) are
called zeros of the system. The poles of the system are the roots of the
characteristic equation. If α is a pole it follows that a(α ) = 0 and that
y(t) = eα t is a solution to the homogeneous equation (3.7). To prove this
we differentiate
dk y
= α k y(t)
dtk
and we find
dn y dn−1 y dn−2 y
+ a 1 + a 2 + . . . + an y = a(α ) y(t) = 0
dtn dtn−1 dtn−2
dn−1u dn−2u
b1 n 1
+ b2 n−2 . . . + bn u = b(β ) Ceβ t = 0.
dt − dt
If the input to the system (3.6) is e β t it thus follows that the solution
is given by the general solution to the homogeneous equation (3.7). The
solution thus will not contain the term e β t. A zero of b(s) at s = β blocks
the transmission of the signal u(t) = Ce β t. Notice that this does not mean
that the output is zero when the the input is e β t unless initial conditions
are chosen in a very special way.
84
3.3 Linear Differential Equations
to an exponential signal. Consider the system given by (3.6) and let the
input be
u(t) = eα t.
The solution to the differential equation is a sum of the general solution to
the homogeneous equation and a particular solution. We will investigate
if there is a particular solution of the form
y(t) = y0 eα t
G (iω ) eiω t = h G (iω )h ei arg G(iω ) eiω t = h G (iω )h ei(ω t+arg G(iω ))
Separating the real and imaginary parts of the input and the output we
find that the input u(t) = sinω t gives the output
X
y(t) = Ck(t) eλ kt + h G (iω )h sin (ω t + arg G (iω )) (3.23)
k
85
Chapter 3. Dynamics
This result is of particular interest for stable systems. For such systems
we have λ k < 0. After an initial transient the response to a sinusoidal
input will thus be sinusoidal with the same frequency as the input. The
output is thus amplified by the factor h G (iω )h and the phase is shifted by
arg G (iω ) in relation to the input. This is discussed further in Section 3.5.
The transform has some properties which makes it very well suited to
deal with linear systems.
First we observe that the transform is linear because
Z ∞ Z ∞
L(a f + bg) = aF (s) + bF (s) = a e−st f (t)dt + b e−stg(t)dt
0 0
Z ∞
= e (a f (t) + bg(t))dt = aL f + bLg
− st
0
(3.25)
Next we show that the Laplace transform of the derivative of a function
is related to the Laplace transform of the function in a very simple way.
If F (s) is the Laplace transform of a function f (t) then the transform of
86
3.4 Laplace Transforms
hence Z t
1 1
L
L f = F (s) f (τ )dτ =
(3.27)
0 s s
Integration of a time function thus corresponds to dividing the Laplace
transform by s. This is consistent with the fact that differentiation of a
time function corresponds to multiplication of the transform with s.
see (3.20). We will now consider the Laplace transform of such an expres-
sion. We have
Z ∞ Z ∞ Z ∞
Y (s) = e−st y(t)dt = e−st g(t − τ )u(τ )dτ dt
0 0 0
Z ∞ Z t
− s( t−τ ) − sτ
= e g(t − τ )u(τ )dτ dt
e
0 0
Z ∞ Z ∞
= e−sτ u(τ )dτ e−stg(t)dt = G (s) U (s)
0 0
87
Chapter 3. Dynamics
The result can be written as Y (s) = G (s) U (s) where G, U and Y are the
Laplace transforms of g, u and y. The system theoretic interpretation is
that the Laplace transform of the output of a linear system is a product
of two terms, the Laplace transform of the input U (s) and the Laplace
transform of the impulse response of the system G (s). A mathematical
interpretation is that the Laplace transform of a convolution is the product
of the transforms of the functions that are convoluted. the fact that the
formula Y (s) = G (s) U (s) is much simpler than a convolution is one reason
why Laplace transforms have become popular in control.
Taking Laplace transforms under the assumption that all initial values
are zero we get.
Hence
b1 sn−1 + b2 sn−2 + . . . + bn−1 s + bn
Y (s) = U (s)
sn + a1 sn−1 + a2 sn−2 + . . . + an−1 s + an
This equation implies that there is a very simple relation between the
Laplace transforms of the inputs and the outputs of a linear system. The
transform of the output is simply the product of the transform of the input
and a rational function G (s), i.e.
where
b1 sn−1 + b2 sn−2 + . . . + bn−1 s + bn
G (s) =
sn + a1 sn−1 + a2 sn−2 + . . . + an−1 s + an
The function G (s) is called the transfer function of the system. For a
single input single output system the transfer function is thus the ratio
of the Laplace transforms of the output and the input where the Laplace
transforms are calculated under the assumption that all initial values
88
3.4 Laplace Transforms
are zero. The fact that all initial values are assumed to be zero has some
consequences that will be discussed later. For a static system we have the
following simple relation between the input u and the output y.
It follows from Equation (3.29) that signals and systems have the same
representations. In the formula we can thus consider g as the input and
u as the transfer function.
Consider the particular case when both the input and the output are
constants, u = u0 and y = y0. Since the derivatives of a constant are zero
it follows from the differential equation (3.28) that
y0 bn
= = G (0)
u0 an
The value of the transfer function for s = 0 thus gives the ratio of the
output and the input in steady state. We call this ratio the steady state
gain.
For a linear static system the input is simply product of the output
and the steady state gain of the system, i.e.
y = gu (3.30)
89
PSfrag replacements
d n
r e u x y
Σ C Σ P Σ
−1
• Use algebra to obtain the transfer functions that relate the signals
of interest.
• Interprete the transfer function.
• Simulate the system by computing responses to interesting signals.
The combination of block diagrams and transfer functions is a powerful be-
cause it is possible both to obtain an overview of a system and find details
of the behavior of the system. By representing signals by their Laplace
transforms and the blocks by transfer functions the relations between sig-
nals in the system are obtained by straight forward manipulations. We
illustrate this by an example.
E = R − Y.
Y = N + X,
X = PV = P( D + U ),
90
3.4 Laplace Transforms
U = CE.
Hence
E = R − N + P( D + CE) = R − N − PD − PCE. (3.31)
1 1 P
E= R− N− D
1 + PC 1 + PC 1 + PC
Hence
1 1 P
E= R− N− D = G er R + G en N + G ed D (3.32)
1 + PC 1 + PC 1 + PC
The error is thus the sum of three terms, depending on the reference r,
the measurement noise n and the load disturbance d. The function
1
G er =
1 + PC
91
Chapter 3. Dynamics
v = Ri
but inductors and capacitors are describe by the linear differential equa-
tions
Z t
Cv = i(τ )dτ
0
di
L =v
dt
V (s) = RI (s)
1
V (s) = I (s)
sC
V (s) = sLI (s)
The transformed equations for all components thus have the form V (s) =
Z (s) I (s), where Z (s) is a generalized impedance, where
92
3.4 Laplace Transforms
PSfrag replacements R C
R1
i
+
V
−
v1 v2
then the voltage V is negligible and the current I0 is zero. The currents
I1 and I2 then are the same which gives
V1(s) V2(s)
=−
Z1 (s) Z2 (s)
93
Chapter 3. Dynamics
Figure 3.8 Computing step responses with better control of the plots.
t=0:0.1:20;
y=step(G,t);
subplot(2,1,1)
ypl=plot(t,y);
set(ypl,’Linewidth’,1.5)
set(gca,’Xtick’,([0 5 10 15 20])
subplot(2,1,2)
plot(t,y,’o’)
The first line defines a vector of times t where the step response is com-
puted. The second line computes the step response for each element of
t. The third line defines a subplot using standard vector notation. The
fourth line defines the plot as the object ylpl. The fifth line changes the
with of the line and the fifth line defines the marking of the time axis.
The sixth line creates a new plot at position 21 and the last line plots the
step response as isolated lines. Notice that all commands can be stored
in a command file and executed as one command. The result is shown in
Figure 3.8.
Matlab can also be used to compute the time functions that correspond to
a Laplace transform as is illustrated in the following example.
94
3.4 Laplace Transforms
Y (s) = G (s)
The Matlab command impulse(Y) thus gives the time function corre-
sponding to the transfer function Y (s).
Matlab is primarily a program for numerical calculations. There are other
programs, Maple and Mathematica for computer algebra. Matlab can how-
ever be used for simple symbolic calculations because it has facilities for
simple manipulation of systems. For example if G and F are linear time
invariant systems models the operations ∗ and / correspond to multiplica-
tion or division of the transfer functions. The following example illustrates
how this can be used.
1
P(s) =
s(s + 1)
1
C(s) = 2 +
s
The step response of the closed loop system can be generated by the fol-
lowing commands.
s=tf([1 0],1) %Define the transfer function s
P=1/(s*(s+2)) %Define process transfer function
C=2+1/s %Define the controller transfer function
T=P*C/(1+P*C) %Compute transfer function from
%reference to output
step(T) %Compute and plot the step response
The transfer functions s is first defined by the command s = tf([10], 1).
The transfer functions for the process and the controller are then defined
in terms of s using a very natural notation. The transfer function from
reference r to output y given by
95
Chapter 3. Dynamics
is then defined and the step response of the closed loop system is then
computed and plotted using the command step(T).
This is perhaps the most natural way to work with transfer functions,
but may be dangerous because Matlab is not smart enough to recognize
cancellations of poles and zeros.
96
3.4 Laplace Transforms
Y (s)
G (s) = = e−sT
U (s)
Vz V2z
=κ 2
Vt V x
where κ = λ /( ρ C) is the thermal diffusivity. To describe how energy is
stored in the system it is necessary to specify the temperature distribu-
tion. The system is therefore and infinite dimensional system. Consider
the particular situation when the input is the temperature at the end of
of the rod and the output is the temperature at a distance a from the end
of an infinitely long rod. Introducing the Laplace transform
Z ∞
Z (s, x) = e−stx(t, x)dt
0
d2 Z (s, x) s
− Z (s, x) = 0
dx 2 κ
97
Chapter 3. Dynamics
Z (s, a) √ √
G (s) = = e−a s/κ = e− sT
U (s)
1
G (s) = √
cosh sT
98
3.5 Frequency Response
b s
F6 (t) = , F7 (t) =
s2 + b2 s2 + b2
Proceeding in this way it is possible to build up tables of transforms that
are useful for hand calculations.
The behavior of the time function for small arguments is governed
by the behavior of the Laplace transform for large arguments. Or more
precisely that the value of f (t) for small t is thus equal to sF (s) for large
s. This is shown as follows.
Z ∞ Z ∞
v
lim sF (s) = lim se−st f (t)dt = lim e−v f ( )dv = f (0)
s→∞ s→∞ 0 s→∞ 0 s
This result, which requires that the limit exists, is called is the initial
value theorem. The converse is also true, we have
Z ∞ Z ∞
v
lim sF (s) = lim se−st f (t)dt = lim e−v f ( )dv = f (∞)
s→0 s→0 0 s→0 0 s
The value of f (t) for large t is thus equal to sF (s) for small s, the result
is called the final value theorem. These properties are very useful for
qualitative assessment of a time functions and Laplace transforms.
99
Chapter 3. Dynamics
If the system is stable, i.e. Reλ k < 0 for all k, the first term will decay
exponentially and the solution will converge to the steady state response
given by
y(t) = h G (iω )h sin (ω t + arg G (iω )) (3.34)
This is illustrated in Figure 3.9 which shows the response of a linear
time-invariant system to a sinusoidal input. The figure shows the output
of the system when it is initially at rest and the steady state output given
by (3.34). The figure shows that after a transient the output is indeed a
sinusoid with the same frequency as the input.
The steady state response to a sinusoid is completely characterized
by the function G (iω ) which is called the frequency response of the sys-
tem. The argument of the function is frequency ω and the function takes
complex values. The magnitude h G (iω )h is called the gain and the an-
gle arg G (iω ) is called the phase. The phase is often negative and the
quantity -arg G (iω ), called the phase lag, is therefore also used. The gain
h G (iω )h is a generalization of the static gain G (0) which tells steady state
output when the input is a constant. It is thus possible to talk about the
gain of the system for signals of different frequencies. The propagation of
any signal can then be obtained by representing it as a sum of sinusoids,
100
3.5 Frequency Response
0.2
0.15
0.1
Output 0.05
−0.05
−0.1
0 5 10 15
0.5
Input
−1
0 5 10 15
Time
investigating each sinusoid individually and adding the outputs using su-
perposition.
The frequency response can be determined experimentally by injecting
a sinusoid and measuring the ratio of the amplitudes and the phase shift
between input and output. Very accurate measurements are possible by
using correlation techniques. This is very important in practice because it
may be very time consuming or even impossible to obtain a mathematical
model from first principles.
101
Chapter 3. Dynamics
Im G ( iω )
Re G ( iω )
g
ϕ
ω
PSfrag replacements
1.4e−s
Figure 3.10 Nyquist plot of the transfer function G ( s) = . The gain and
( s + 1)2
phase for the frequency ω are g = h G ( iω )h and ϕ = arg G ( iω ) .
B A
L( s)
PSfrag replacements
−1
and a controller with transfer function C(s). Introduce the loop transfer
function
L(s) = P(s) C(s), (3.35)
which represents signal transmission around the loop. The system can
then be represented by the block diagram in Figure 3.11. We will first
determine conditions for having a periodic oscillation in the loop. Assume
102
3.5 Frequency Response
that the feedback loop is cut as indicated in the figure and that a sinusoid
of frequency ω is injected at point A. In steady state the signal at point B
will also be a sinusoid with the same frequency. It seems reasonable that
an oscillation can be maintained if the signal at B has the same amplitude
and phase as the injected signal because we could then connect A to B.
Tracing signals around the loop we find that the condition that the signal
at B is identical to the signal at A is that
L(iω 0 ) = −1 (3.36)
This condition has a nice interpretation in the Nyquist plot. It means that
the Nyquist plot of L(iω ) intersects the negative real axis at the point -
1. The frequency where the intersection occurs is the frequency of the
oscillation.
Intuitively it seems reasonable that the system would be stable if the
critical point -1 is on the left hand side of the Nyquist curve as indicated
in Figure 3.11. This means that the signal at point B will have smaller
amplitude than the injected signal. This is essentially true, but there are
several subtleties, that requires a proper mathematics to clear up. This
will be done later. The precise statement is given by Nyquist’s stability
criterion.
Stability Margins
In practice it is not enough that the system is stable. There must also be
some margins of stability. There are many ways to express this. Many of
the criteria are inspired by Nyquist’s stability criterion. They are based
on the fact that it is easy to see the effects of changes of the gain and
the phase of the controller in the Nyquist diagram of the loop transfer
function L(s). An increase of controller gain simply expands the Nyquist
plot radially. An increase of the phase of the controller twists the Nyquist
plot clockwise, see Figure 3.12. The gain margin g m tells how much the
controller gain can be increased before reaching the stability limit. Let
ω 180 be the smallest frequency where the phase lag of the loop transfer
function L(s) is 180○. The gain margin is defined as
1
gm = (3.37)
h L(iω 180 )h
103
Chapter 3. Dynamics
Im L( iω )
−1 −1/gm Re L( iω )
PSfrag replacements ϕm
Figure 3.12 Nyquist plot of the loop transfer function L with gain margin g m ,
phase margin ϕ m and stability margin d.
The phase margin ϕ m is the amount of phase lag required to reach the
stability limit. Let ω g c denote the lowest frequency where the loop transfer
function L(s) has unit magnitude. The phase margin is then given by
104
3.5 Frequency Response
0.5
400
0.4
300
0.3
200
0.2
100 0.1
0 0
−0.1
−100
−0.2
−200
−0.3
−300
−0.4
−400 −0.5
−400 −300 −200 −100 0 100 200 300 400 −9 −8 −7 −6 −5 −4 −3 −2 −1 0 1
3(s+1)2
Figure 3.13 Nyquist curve for the loop transfer function L( s) = s(s+6)2
. The plot
on the right is an enlargement of the area around the origin of the plot on the left.
then necessary to consider the intersections that are closest to the critical
point. For more complicated systems there is also another number that is
highly relevant namely the delay margin. The delay margin is defined as
the smallest time delay required to make the system unstable. For loop
transfer functions that decay quickly the delay margin is closely related
to the phase margin but for systems where the amplitude ratio of the loop
transfer function has several peaks at high frequencies the delay margin
is a much more relevant measure.
3(s + 1)2
L(s) = (3.39)
s(s + 6)2
The Nyquist plot of the loop transfer function is shown in Figure 3.13.
Notice that the Nyquist curve intersects the negative real axis twice. The
first intersection occurs at s = − for ω = and the second at s = − for
ω =. The intuitive argument based on signal tracing around the loop is
less intuitive in this case, because injection of a sinusoid with frequency
105
Chapter 3. Dynamics
PROOF 3.1
Assume that z = a is a zero of multiplicity m. In the neighborhood of
z = a we have
f ( z) = ( z − a) m g( z)
where the function g is analytic and different form zero. We have
f P ( z) m gP ( z)
= +
f ( z) z− a g( z)
d f P ( z)
log f ( z) =
dz f ( z)
106
3.5 Frequency Response
∆ Γ log f ( z) = i∆ Γ arg f ( z)
REMARK 3.1
The number wn is called the winding number.
REMARK 3.2
The theorem is useful to determine the number of poles and zeros of
an function of complex variables in a given region. To use the result we
must determine the winding number. One way to do this is to investigate
how the curve Γ is transformed under the map f . The variation of the
argument is the number of times the map of Γ winds around the origin
in the f -plane. This explains why the variation of the argument is also
called the winding number.
We will now use the Theorem 1 to prove Nyquist’s stability theorem.
For that purpose we introduce a contour that encloses the right half plane.
For that purpose we choose the contour shown in Figure 3.14. The contour
consists of a small half circle to the right of the origin, the imaginary axis
and a large half circle to the right with with the imaginary axis as a
diameter. To illustrate the contour we have shown it drawn with a small
radius r and a large radius R. The Nyquist curve is normally the map
of the positive imaginary axis. We call the contour Γ the full Nyquist
contour.
Consider a closed loop system with the loop transfer function L(s).
The closed loop poles are the zeros of the function
f (s) = 1 + L(s)
To find the number of zeros in the right half plane we thus have to inves-
tigate the winding number of the function f = 1 + L as s moves along the
contour Γ . The winding number can conveniently be determined from the
Nyquist plot. A direct application of the Theorem 1 gives.
107
Chapter 3. Dynamics
k
L(s) =
s(s + 1)2
Figure 3.15 shows the image of the contour Γ under the map L. The
Nyquist plot intersects the imaginary axis for ω = 1 the intersection is
at − k/2. It follows from Figure 3.15 that the winding number is zero if
k < 2 and 2 if k > 2. We can thus conclude that the closed loop system is
stable if k < 2 and that the closed loop system has two roots in the right
half plane if k > 2.
By using Nyquist’s theorem it was possible to resolve a problem that
had puzzled the engineers working with feedback amplifiers. The follow-
ing quote by Nyquist gives an interesting perspective.
Mr. Black proposed a negative feedback repeater and proved by tests
that it possessed the advantages which he had predicted for it. In
108
3.5 Frequency Response
k
Figure 3.15 Map of the contour Γ under the map L( s) = s((s+1)2
. The curve is
drawn for k < 2. The map of the positive imaginary axis is shown in full lines, the
map of the negative imaginary axis and the small semi circle at the origin in dashed
lines.
particular, its gain was constant to a high degree, and it was lin-
ear enough so that spurious signals caused by the interaction of the
various channels could be kept within permissible limits. For best re-
sults, the feedback factor, the quantity usually known as µ β (the loop
transfer function), had to be numerically much larger than unity. The
possibility of stability with a feedback factor greater than unity was
puzzling. Granted that the factor is negative it was not obvious how
that would help. If the factor was -10, the effect of one round trip
around the feedback loop is to change the magnitude of the current
from, say 1 to -10. After a second trip around the loop the current
becomes 100, and so forth. The totality looks much like a divergent
series and it was not clear how such a succession of ever-increasing
components could add to something finite and so stable as experience
had shown. The missing part in this argument is that the numbers
that describe the successive components 1, -10, 100, and so on, rep-
resent the steady state, whereas at any finite time many of the com-
ponents have not yet reached steady state and some of them, which
are destined to become very large, have barely reached perceptible
magnitude. My calculations were principally concerned with replac-
ing the indefinite diverging series referred to by a series which gives
the actual value attained at a specific time t. The series thus obtained
is convergent instead of divergent and, moreover, converges to values
in agreement with the experimental findings.
109
Chapter 3. Dynamics
Figure 3.16 Map of the contour Γ under the map L( s) = s(s−1k)(s+5) . The curve on
the right shows the region around the origin in larger scale. The map of the positive
imaginary axis is shown in full lines, the map of the negative imaginary axis and
the small semi circle at the origin in dashed lines.
Notice that Nyquist’s theorem does not hold if the loop transfer func-
tion has a pole in the right half plane. There are extensions of the Nyquist
theorem to cover this case but it is simpler to invoke Theorem 1 directly.
We illustrate this by two examples.
110
3.5 Frequency Response
Figure 3.17 Map of the contour Γ under the map L( s) = ss2+−21 given by (3.40). The
map of the positive imaginary axis is shown in full lines, the map of the negative
imaginary axis and the small semi circle at the origin in dashed lines.
Since the loop transfer function has a pole in the right half plane we have
P = 1 and we get N = 2. The characteristic equation thus has two roots
in the right half plane.
s+2
L(s) = (3.40)
s2 − 1
This transfer function has one pole at s = 1 in the right half plane.
The Nyquist plot of the loop transfer function is shown in Figure 3.17.
Traversing the contour Γ in clockwise we find that the winding number
is -1. Applying Theorem 1 we find that
N − P = −1
Since the loop transfer function has a pole in the right half plane we have
P = 1 and we get N = 0. The characteristic equation thus has no roots
in the right half plane and the closed loop system is stable.
Bode Plots
he Nyquist plot is one way to represent the frequency response G (iω ).
Another useful representation was proposed by Bode who represented it
by two curves, the gain curve and the phase curve. The gain curve gives
gain h G (iω )h as a function of ω and the phase curve phase arg G (iω ) as
a function of ω . The curves are plotted as shown below with logarithmic
111
Chapter 3. Dynamics
3
10
h G ( iω )h 2
10
1
10
−2 −1 0 1 2
10 10 10 10 10
100
arg G ( iω )
50
−100
−2 −1 0 1 2
10 10 10 10 10
ω
Figure 3.18 Bode plot of the transfer function of the ideal PID controller C ( s) =
20 + 10/ s + 10s. The top plot is the gain curve and bottom plot is the phase curve.
The dashed lines show straight line approximations of the curves.
scales for frequency and magnitude and linear scale for phase, see Fig-
ure 3.18 An useful feature of the Bode plot is that both the gain curve and
the phase curve can be approximated by straight lines, see Figure 3.18
where the approximation is shown in dashed lines. This fact was partic-
ularly useful when computing tools were not easily accessible. The fact
that logarithmic scales were used also simplified the plotting.
It is easy to sketch Bode plots because with the right scales they have
linear asymptotes. This is useful in order to get a quick estimate of the
behavior of a system. It is also a good way to check numerical calculations.
Consider first a transfer function which is a polynomial G (s) = b(s)/a(s).
We have
log G (s) = log b(s) − log a(s)
Since a polynomial is a product of terms of the type :
s, s + a, s2 + 2ζ as + a2
it suffices to be able to sketch Bode diagrams for these terms. The Bode
plot of a complex system is then obtained by composition.
G (s) = s
112
3.5 Frequency Response
1
10
h G ( iω )h
0
10
−1
10
−1 0 1
10 10 10
91
arg G ( iω )
90.5
90
89
−1 0 1
10 10 10
ω
The gain curve is thus a straight line with slope 1 and the phase curve is
a constant at 90○.. The Bode plot is shown in Figure 3.19
The gain curve is thus a straight line with slope -1 and the phase curve
is a constant at −90○. The Bode plot is shown in Figure 3.20
Compare the Bode plots for the differentiator in Figure 3.19 and the
integrator in Figure 3.20. The sign of the phase is reversed and the gain
curve is mirror imaged in the horizontal axis. This is a consequence of
the property of the logarithm.
1
log = − log G = − log h G h − i arg G
G
113
Chapter 3. Dynamics
1
10
h G ( iω )h
0
10
−1
10
−1 0 1
10 10 10
−89
−89.5
arg G ( iω )
−90
−91
−1 0 1
10 10 10
ω
G (s) = s + a
We have
G (iω ) = a + iω
p
h G (iω )h = ω 2 + a2 , arg G (iω ) = arctan ω /a
Hence
1
log h G (iω )h = log (ω 2 + a2 ), arg G (iω ) = arctan ω /a
2
The Bode Plot is shown in Figure 3.21. Both the gain curve and the phase
curve can be approximated by straight lines if proper scales are chosen
114
3.5 Frequency Response
1
10
h G ( iω )h/ a
0
10
−1 0 1
10 10 10
100
arg G ( iω )
50
PSfrag replacements
0
−1 0 1
10 10 10
ω /a
Figure 3.21 Bode plot of the transfer function G ( s) = s + a. The dashed lines
show the piece-wise linear approximations of the curves.
Notice that a first order system behaves like an integrator for high fre-
quencies. Compare with the Bode plot in Figure 3.20.
G (s) = s2 + 2aζ s + a2
We have
G (iω ) = a2 − ω 2 + 2iζ aω
115
Chapter 3. Dynamics
Hence
1
log h G (iω )h = log (ω 4 + 2a2ω 2 (2ζ 2 − 1) + a4 )
2
arg G (iω ) = arctan 2ζ aω /(a2 − ω 2 )
Notice that the smallest value of the magnitude minω h G (iω )h = 1/2ζ is
obtained for ω = a The gain is thus constant for small ω . It has an asymp-
tote with zero slope for low frequencies. For large values of ω the gain is
proportional to ω 2 , which means that the gain curve has an asymptote
with slope 2. The phase is zero for low frequencies and approaches 180○
for large frequencies. The curves can be approximated with the following
piece-wise linear expressions
2 log a if ω << a,
log h G (iω )h 2 log a + log 2ζ if ω = a, ,
2 log ω if ω >> a
0 if ω << a,
π ω −a
arg G (iω ) + if ω = a, ,
2 aζ
π if ω >> a
The Bode Plot is shown in Figure 3.22, the piece-wise linear approxima-
tions are shown in dashed lines.
116
3.5 Frequency Response
2
10
h G ( iω )h/ a
10
1
0
10
−1
10
−1 0 1
10 10 10
200
arg G ( iω )
150
100
PSfrag replacements 50
0
−1 0 1
10 10 10
ω /a
200(s + 1) 1+s
G (s) = =
s(s + 10)(s + 200) 10s(1 + 0.1s)(1 + 0.01s)
The break points are 0.01, 0.1, 1. For low frequencies the transfer function
can be approximated by
1
G (s)
10s
Following the procedure we get
• The low frequencies the system behaves like an integrator with gain
0.1. The low frequency asymptote thus has slope -1 and it crosses
the axis of unit gain at ω = 0.1.
• The first break point occurs at ω = 0.01. This break point corre-
sponds to a pole which means that the slope decreases by one unit
to -2 at that frequency.
• The next break point is at ω = 0.1 this is also a pole which means
that the slope decreases to -3.
117
Chapter 3. Dynamics
0
Gain
10
−1
10
−2
10
−3
10
−1 0 1 2 3
10 10 10 10 10
ω
Phase
0
−50
−100
PSfrag replacements
−150
h G ( iω )h 10
−1
10
0
10
1 2
10 10
3
arg G ( iω ) ω
Figure 3.23 Illustrates how the asymptotes of the gain curve of the Bode plot can
be sketched. The dashed curves show the asymptotes and the full lines the complete
plot.
118
3.5 Frequency Response
gain
1
10
0
10
−1
10
−1 0
10 10
phase
−100
−120
−140
PSfrag replacements
−160
Gain
Phase −180
ω
h G ( iω )h −200
arg G ( iω )
−1 0
10 10
Figure 3.24 Finding gain and phase margins from the Bode plot of the loop trans-
fer function.
119
Chapter 3. Dynamics
Bode’s Relations
Analyzing the Bode plots in the examples we find that there appears
to be a relation between the gain curve and the phase curve. Consider
e.g. the curves for the differentiator in Figure 3.19 and the integrator
in Figure 3.20. For the differentiator the slope is +1 and the phase is
constant pi/2 radians. For the integrator the slope is -1 and the phase
is − pi/2. Bode investigated the relations between the curves and found
that there was a unique relation between amplitude and phase for many
systems. In particular he found the following relations for system with no
poles and zeros in the right half plane.
Z
2ω 0 ∞ log h G (iω )h − log h G (iω 0 )h
arg G (iω 0 ) = dω
π 0 ω 2 − ω 20
Z
1 ∞ d log h G (iω )h ω + ω0
= log d log ω
π 0 d log ω ω − ω0
Z ∞
π ω d log h G (iω )h
= w d log ω
2 0 ω0 d log ω
π d log h G (iω )h (3.41)
2 d log ω
Z
log h G (iω )h 2ω 20 ∞ ω −1 arg G (iω ) − ω − 0 arg G ( iω 0 )
1
=− dω
log h G (iω 0 )h π 0 ω 2 − ω 20
Z
2ω 2 ∞ d ω −1 arg G (iω ) ω + ω0
=− 0 log dω
π 0 dω ω − ω0
See Figure 3.25. The formula for the phase tells that the phase is a
weighted average of the logarithmic derivative of the gain, approxima-
tively
π d log h G (iω )h
arg G (iω ) (3.42)
2 d log ω
This formula implies that a slope of +1 corresponds to a phase of π /2,
which holds exactly for the differentiator, see Figure 3.19.
120
3.5 Frequency Response
4
w
PSfrag replacements 1
0
−3 −2 −1 0 1 2 3
10 10 10 10 10 10 10
ω /ω 0
Figure 3.25 The weighting kernel in Bodes formula for computing the phase from
the gain.
nice property of these systems is that the phase curve is uniquely given
by the gain curve. These systems are also relatively easy to control. Other
systems have larger phase lag, i.e. more negative phase. These systems
are said to be non-minimum phase, because they have more phase lag than
the equivalent minimum phase systems. Systems which are not minimum
phase are more difficult to control. Before proceeding we will give some
examples.
121
Chapter 3. Dynamics
1
Gain
10
0
10
−1
10
−1 0 1
10 10 10
ω
Phase
0
−100
−300
h G ( iω )h
−400
−1 0 1
10 10 10
arg G ( iω ) ω
Figure 3.26 Bode plot of a time delay which has the transfer function G ( s) = e −s .
Notice that the gain is one. The minimum phase system which has unit
gain has the transfer function G (s) = 1. In Figure 3.27 we show the Bode
plot of the transfer function. The Bode plot resembles the Bode plot for a
time delay which is not surprising because the exponential function e−sT
can be approximated by
1 − sT /2
e−sT =
1 + sT /2
The largest phase lag of a system with a zero in the RHP is however
pi.
We will later show that the presence of a zero in the right half plane
severely limits the performance that can be achieved. We can get an intu-
itive feel for this by considering the step response of a system with a right
122
3.5 Frequency Response
1
10
Gain h G ( iω )h
0
10
−1
10
−1 0 1
10 10 10
Phase arg G ( iω )
0
PSfrag replacements
−50
−100
h G ( iω )h −150
−1 0 1
10 10 10
arg G ( iω ) ω
a−s
Figure 3.27 Bode plot of the transfer function G ( s) = The phase of the
a+s
corresponding minimum phase system G ( s) = 1 is shown in dashed lines.
half plane zero. Consider a system with the transfer function G (s) that
has a zero at s = −α in the right half plane. Let h be the step response
of the system. The Laplace transform of the step response is given by
Z t
G (s)
H (s) = = e−sth(t)dt
s 0
123
Chapter 3. Dynamics
0.5
PSfrag replacements
Gain 0
Phase
ω
h G ( iω )h −0.5
Figure 3.28 Step response of a system with a zero in the right half plane. The
6(− s + 1)
system has the transfer function G ( s) = .
s2 + 5s + 6
124
3.5 Frequency Response
1
Gain
10
0
10
−1
10
−1 0 1
10 10 10
ω
Phase
0
−50
PSfrag replacements
−100
−150
h G ( iω )h
−1 0 1
10 10 10
arg G ( iω ) ω
s+a
Figure 3.29 Bode plot of a the transfer function G ( s) = which has a pole
s−a
in the right half plane.
125
Chapter 3. Dynamics
dx
= f ( x, u)
dt (3.43)
y = g( x, u)
The dimension of the state vector is called the order of the system.
The system is called time-invariant because the functions f and g do not
depend explicitly on time t. It is possible to have more general time-
varying systems where the functions do depend on time. The model thus
consists of two functions. The function f gives the velocity of the state
vector as a function of state x, control u and time t and the function g
gives the measured values as functions of state x, control u and time t. The
function f is called the velocity function and the function g is called the
sensor function or the measurement function. A system is called linear
if the functions f and g are linear in x and u. A linear system can thus
be represented by
dx
= Ax + Bu
dt
y = Cx + Du
where A, B, C and D are constant varying matrices. Such a system is
said to be linear and time-invariant, or LTI for short. The matrix A is
126
3.6 State Models
PSfrag replacements
l
θ x
called the dynamics matrix, the matrix B is called the control matrix, the
matrix C is called the sensor matrix and the matrix D is called the direct
term. Frequently systems will not have a direct term indicating that the
control signal does not influence the output directly. We will illustrate
by a few examples.
d2θ
J = mg l sin θ + mul cos θ
dt2
127
Chapter 3. Dynamics
R L +
M J ω
u M e
− D
i
(a) (b)
dω
J + Dω = kI
dt
128
3.6 State Models
dI dω
E = RI + L +V−k
dt dt
dV
I=C
dt
This is a linear time-invariant system with three state variables and one
input.
dV
= qin − qout
dt
129
Chapter 3. Dynamics
Equilibria
To investigate a system we will first determine the equilibria. Consider
the system given by (3.43) which is assumed to be time-invariant. Let the
control signal be constant u = u0 . The equilibria are states x0 such that
the dx/dt = 0. Hence
f ( x0 , u 0 ) = 0
Notice that there may be several equilibria.
For second order systems the state equations can be visualized by plot-
ting the velocities for all points in the state space. This graph is called the
phase plane shows the behavior qualitative. The equilibria corresponds to
points where the velocity is zero. We illustrate this with an example.
130
3.6 State Models
3
x’=y−y
2
y’=−x−y
1.5
0.5
0
y
−0.5
−1
−1.5
−2
Figure 3.33 Phase plane for the second order system dx1 / dt = x2 − x23 ,dx2 / dt =
− x1 − x22 .
Linearization
Nonlinear systems are unfortunately difficult. It is fortunate that many
aspects of control can be understood from linear models. This is particu-
larly true for regulation problems where it is intended to keep variables
close to specified values. When deviations are small the nonlinearities can
be approximated by linear functions. With efficient control the deviations
are small and the approximation works even better. In this section we will
show how nonlinear dynamics systems are approximated. We will start
with an example that shows how static systems are approximated.
y − y0 = gP (u0)(u − u0 )
The linearized model thus replaces the nonlinear curve by its tangent at
the operating point.
131
Chapter 3. Dynamics
dx
= f ( x, u)
dt
y = g( x, u)
x = x0 + δ x , u = u0 + δ u, y = y0 + δ y
dx Vf Vf
= f ( x0 + δ x, u0 + δ u) f ( x0, u0 ) + ( x0 , u0 )δ x + ( x0 , u0 )δ u
dt Vx Vu
Vg Vg
y = g( x0 + δ x, u0 + δ u) y0 + ( x0 , u0 )δ x + ( x0, u0 )δ u
Vx Vu
We have f ( x0, u0 ) = 0 because x0 is an equilibrium and we find the fol-
lowing approximation for small deviations around the equilibrium.
d( x − x0 )
= A( x − x0 ) + B (u − u0 )
dt
y − y0 = C( x − x0 ) + D (u − u0 )
where
Vf Vf
A= ( x0 , u 0 ) B= ( x0 , u 0 )
Vx Vu
Vg Vg
C= ( x0 , u 0 ) D= ( x0 , u 0 )
Vx Vu
The linearized equation is thus a linear time-invariant system, compare
with (3.48). It is common practice to relabel variables and simply let x, y
and u denote deviations from the equilibrium.
We illustrate with a few examples
132
3.6 State Models
133
Chapter 3. Dynamics
dx
= x2
dt
x(0) = 1
1
x(t) =
1−t
This example illustrates that solutions can blow up in finite time. The rea-
son for this is that since the velocity is proportional to x 2 it increases very
rapidly with increasing x. This type of behavior cannot occur for linear
systems because the solutions to such systems are composed of polyno-
mials and exponential functions. The technical name for this behavior is
finite escape time.
The next example shown that there may be problems with uniqueness
even of simple differential equations.
134
3.6 State Models
100
80
60
x
40
20
PSfrag replacements
0
0 1 2 3 4 5 6 7 8 9 10
t
dx √
=2 x
dt (3.47)
x(0) = 0
h f ( x) − f ( y)h ≤ Ch x − yh
where C is a constant. The square root function does not have this prop-
erty for y = 0 because its derivative is infinite.
Notice that the water tank in Example 3.28 has a model where the
velocity depends on the square root of the height h. We could then expect
that difficulties may occur in this case. Notice, however, that the physical
model is no longer valid if the height is so small so that the water level
is below the upper part of the outlet area.
135
Chapter 3. Dynamics
The model
dx
= Ax + Bu
dt (3.48)
y = Cx + Du
is one of the standard models in control. In this section we will present
an in depth treatment. Let us first recall that x is the state vector, u the
control, y the measurement. The model is nice because it can represent
systems with many inputs and many outputs in a very compact form.
Because of the advances in numeric linear algebra there are also much
powerful software for making computations. Before going into details we
will present some useful results about matrix functions. It is assumed
that the reader is familiar with the basic properties of matrices.
Matrix Functions
Some basic facts about matrix functions are summarized in this section.
Let A be a square matrix, since it is possible to compute powers of matrices
we can define a matrix polynomial as follows
f ( A) = a0 I + a1 A + . . . + a n An
f ( A) = a0 I + a1 A + . . . + a n An + . . .
de At 1 1
= A + A 2 t + A3 t2 + . . . + An tn−1 + . . .
dt 2 (n − 1)!
1 1
= A(= I + At + ( At)2 + . . . + An tn + . . .) = Ae At
2 n!
The matrix exponential thus has the property
de At
= Ae At = e At A (3.49)
dt
Matrix functions do however have other interesting properties. One result
is the following.
136
3.7 Linear Time-Invariant Systems
PROOF 3.2
If a matrix has distinct eigenvalues it can be diagonalized and we have
A = T −1 Λ T. This implies that
A2 = T −1 Λ T T −1Λ T = T −1 Λ 2 T
A3 = T −1 Λ T A2 = T −1 Λ T T −1Λ 2 T = T −1Λ 3 T
λ ni + a1 λ ni −1 + a2 λ ni −2 . . . + an = 0
Hence
Λ ni + a1 Λ ni −1 + a2 Λ ni −2 . . . + an I = 0
Multiplying by T −1 from the left and T from the right and using the
relation Ak = T −1 Λ k T now gives
An + a1 An−1 + a2 An−2 . . . + an I = 0
f ( A) = c0 I + c1 A + . . . + c k−1 Ak−1
137
Chapter 3. Dynamics
Z t
x(t) = e At x(0) + e A(t−τ ) Bu(τ )dτ (3.50)
0
To prove this we differentiate both sides and use the property (3.49) of
the matrix exponential. This gives
Z t
dx
= Ae At x(0) + Ae A(t−τ ) Bu(τ )dτ + Bu(t) = Ax + Bu
dt 0
which prove the result. Notice that the calculation is essentially the same
as for proving the result for a first order equation.
Input-Output Relations
It follows from Equations (3.48) and (3.50) that the input output relation
is given by
Z t
y(t) = Ce At x(0) + e A(t−τ ) Bu(τ )dτ + Du(t)
0
Taking the Laplace transform of (3.48) under the assumption that x(0) =
0 gives
Solving the first equation for X (s) and inserting in the second gives
138
3.7 Linear Time-Invariant Systems
1 s 1 −1 0 1
−1 1
G (s) = C[sI − A] B= 2 0
= 2
s −1 1 s 1 s −1
To find the input output relation we can differentiate the output and we
obtain
y = Cx
dy dx
=C = CAx + CBu
dt dt
d2 y dx du du
= CA + CB = CA2 x + CABu + CB
dt2 dt dt dt
..
.
dn y n n−1 n−2 du dn−1u
= CA x + CA Bu + CA B + . . . + CB
dtn dt dtn−1
Let ak be the coefficients of the characteristic equation. Multiplying the
first equation by a n , the second by a n−1 etc we find that the input-output
relation can be written as.
dn y dn−1 y dn−1u dn−2u
+ a 1 + . . . + a n y = B 1 + B 2 + . . . + Bnu,
dtn dtn−1 dtn−1 dtn−2
139
Chapter 3. Dynamics
B1 = CB
B2 = CAB + a1 CB
B3 = CA2 B + a1 CAB + a2 CB
..
.
Bn = CAn−1 B + a1 CAn−1 B + . . . + a n−1 CB
This equation has a solution with nonzero x0 , u0 only if the matrix on the
left has nonzero rank. The zeros are thus the values s such that
sI − A B
det
=0 (3.52)
C D
Notice in particular that if the matrix B has full rank the matrix has
n linearly independent rows for all values of s. Similarly there are n
linearly independent columns if the matrix C has full rank. This implies
that systems where the matrices B or C are of full rank do not have
zeros. In particular it means that a system has no zeros if the full state
is measured.
140
3.7 Linear Time-Invariant Systems
dz
= T ( Ax + Bu) = T AT −1 z + T Bu = Ãz + B̃u
dt
y = Cx + DU = CT −1 z + Du = C̃z + Du
The transformed system has the same form as (3.48) but the matrices A,
B and C are different
−1 t
g̃(t) = C̃e Ãt B̃ = CT −1 eT AT T B = Ce At B = g(t)
Canonical Forms
There are special choices of coordinates that give state equations with a
special structure.
141
Chapter 3. Dynamics
λ1 0 β1
dz
λ2
β2
=
z +
u
dt . .
.
.
.
.
(3.54)
0 λn βn
y = γ 1 γ 2 . . . γ n z + Du
n
X β iγ i
G (s) = +D
s − λi
i=1
dn y dn−1 y
n
+ a1 n−1 + . . . + an y = u (3.55)
dt dt
1
G phc(s) = (3.56)
sn + a1 sn−1 + . . . + an−1 s + an
The system is a special case of an LTI system with one input and one
output. Introduce the state variables as the output and its derivatives,
i.e.
dn−1 y dn−2 y dy
x1 = , x2 = , ⋅ ⋅ ⋅ , xn−1 = , xn = y
dtn−1 dtn−2 dt
142
3.7 Linear Time-Invariant Systems
It follows that
dx1
= − a 1 x1 − a 2 x2 − ⋅ ⋅ ⋅ − a n x n + u
dt
dx2
= x1
dt
dx3
= x2
dt
..
.
dxn
= xn−1
dt
or in matrix form
−a1 −a2 . . . an−1 −an 1
1 0 0 0
0
dx
0 1 0 0
0
=
x +
u
dt
..
. (3.57)
.
.
.
0 0 1 0 0
y = 0 0 ... 0 1 x
Expanding the determinant by the last row we find that the following
recursive equation for the polynomial D n (s).
Dn (s) = sDn−1(s) + an
143
Chapter 3. Dynamics
dn y dn−1 y dn−1u
n
+ a1 n−1 + . . . + an y = b1 n−1 + . . . + bn u
dt dt dt
where x1 , x2 , etc are given by (3.57). It thus follows that the state equa-
tions can be written as
−a1 −a2 . . . an−1 −an 1
1 0 0 0
0
dz
0 1 0 0
0
=
z +
u
dt
.
.
(3.59)
..
..
0 0 1 0 0
y = b1 b2 . . . bn−1 bn z + Du
b1 sn−1 + b2 sn−2 + . . . + bn
G (s) = +D
sn + a1 sn−1 + a2 sn−2 + . . . + an
144
3.7 Linear Time-Invariant Systems
b1 sn−1 + b2 sn−2 + . . . + bn
G (s) =
sn + a1 sn−1 + a2 sn−2 + . . . + an
b1 sn−1 + b2 sn−2 + . . . + bn
X1 = Y = U
sn + a1 sn−1 + a2 sn−2 + . . . + an
then
sX 1 = −a1 X 1 + b1 U + X 2
where
sn−1 X 2 = −(a2 sn−2 + a3 sn−3 + . . . + an X 1
+ b2 sn−2 + b3 sn−3 + . . . + bn ) U
sX 2 = −a2 X 2 + b2 U + X 3
where
sn−2 X 3 = −(a3 sn−3 + a4n−4 . . . + an X 1 + b3 sn−3 + . . . + bn ) U
sX 3 = −a3 X 1?b3 U + X 4
X n = −an X 1 + b1 U
145
Chapter 3. Dynamics
Collecting the different parts and converting to the time domain we find
that the system can be written as
− a1 1 0 ... 0 b1
0 1 0 b2
− a2
dz
.
.
=
..
z +
.
.
u
dt
(3.60)
−an−1 0 0 1
bn−1
−an 0 0 0 bn
y = 1 0 0... 0 z + Du
b1 sn−1 + b2 sn−2 + . . . + bn
G (s) = +D
sn + a1 sn−1 + a2 sn−2 + . . . + an
Reachability
We will now disregard the measurements and focus on the evolution of
the state which is given by
dx
= Ax + Bu
dt
where the system is assumed to be or order n. A fundamental question
is if it is possible to find control signals so that any point in the state
space can be reached. For simplicity we assume that the initial state of
the system is zero.
We will first provide an heuristic argument based on formal calcula-
tions with delta functions. When the initial state is zero The response of
the state to a unit step in the input is given by
Z t
x(t) = e A(t−τ ) B dτ = A−1 ( e At − I ) B (3.61)
0
The derivative of a unit step function is the delta function δ (t) which may
be regarded as a function which is zero everywhere except at the origin
with the property that Z ∞
δ (t)dt = 1
∞
The response of the system to a delta function is thus given by the deriva-
tive of (3.61)
dx
= e At B
dt
146
3.7 Linear Time-Invariant Systems
dx
= Ae At B
dt
The input
x(t) = α 1 e At B + α 2 Ae At B + α 3 A2 e At B + ⋅ ⋅ ⋅ + α n An−1 e At B
Hence
x(0+) = α 1 B + α 2 AB + α 3 A2 B + ⋅ ⋅ ⋅ + α n An−1 B
The right hand is a linear combination of the columns of the matrix.
Wr = B AB . . . A n−1 B (3.62)
Again we observe that the right hand side is a linear combination of the
columns of the reachability matrix Wr given by (3.62).
We illustrate by two examples.
147
Chapter 3. Dynamics
This matrix has full rank and we can conclude that the system is reach-
able.
where
w0 = B̃
w1 = a1 B̃ + Ã B̃
..
.
wn−1 = an−1 B + an−2 ÃB + ⋅ ⋅ ⋅ + Ãn−1 B
148
3.7 Linear Time-Invariant Systems
S
u
wk = ak + w̃k−1
which shows that the matrix (3.64) is indeed the inverse of W̃r .
dx1
= − x1 + u
dt
dx2
= − x2 + u
dt
149
Chapter 3. Dynamics
Coordinate Changes
It is interesting to investigate how the reachability matrix transforms
when the coordinates are changed. Consider the system in (3.48). Assume
that the coordinates are changed to z = T x. It follows from (3.53) that
the dynamics matrix and the control matrix for the transformed system
are
à = T AT −1
B̃ = T B
We have
à B̃ = T AT −1 T B = T AB
Ã2 B̃ = (T AT −1)2 T B = T AT −1 T AT −1 T B = T A2 B
..
.
Ãn B̃ = T An B
and we find that the reachability matrix for the transformed system has
the property
W̃r = B̃ Ã B̃ . . . Ãn−1 B̃ = T B AB . . . A n−1 B = T Wr
(3.65)
This formula is very useful for finding the transformation matrix T.
Observability
When discussing reachability we neglected the output and focused on the
state. We will now discuss a related problem where we will neglect the
input and instead focus on the output. Consider the system
dx
= Ax
dt (3.66)
y = Cx
150
3.7 Linear Time-Invariant Systems
151
Chapter 3. Dynamics
S
y
Σ
which has full rank. It is thus possible to compute the state from a mea-
surement of the angle.
A Non-observable System
It is useful to have an understanding of the mechanisms that make a
system unobservable. Such a system is shown in Figure 3.36. Next we
will consider the system in (3.60) on observable canonical form, i.e.
− a1 1 0 ... 0 b1
− a 2 0 1 0
b2
dz
..
..
= .
z +
u
dt
.
− a n−1 0 0 1
bn−1
−an 0 0 0 bn
y = 1 0 0... 0 z + Du
A straight forward but tedious calculation shows that the inverse of the
observability matrix has a simple form. It is given by
1 0 0 ... 0
a 1 1 0 ... 0
Wo−1 a a1 1 ... 0
2
=
..
.
an−1 an−2 an−3 ... 1
152
3.7 Linear Time-Invariant Systems
Coordinate Changes
It is interesting to investigate how the observability matrix transforms
when the coordinates are changed. Consider the system in (3.48). Assume
that the coordinates are changed to z = T x. It follows from (3.53) that
the dynamics matrix and the output matrix are given by
à = T AT −1
C̃ = CT −1
We have
C̃ Ã = CT −1 T AT −1 = CAT −1
C̃ Ã2 = CT −1(T AT −1)2 = CT −1 T AT −1 T AT −1 = CA2 T −1
..
.
C̃ Ãn = CAn T −1
and we find that the observability matrix for the transformed system has
the property
C̃
C̃ Ã
C̃ Ã2 −1
W̃o =
T = Wo T −1 (3.68)
.
.
.
n−1
C̃ Ã
This formula is very useful for finding the transformation matrix T.
153
Chapter 3. Dynamics
Kalman’s Decomposition
The concepts of reachability and observability make it possible understand
the structure of a linear system. We first observe that the reachable states
form a linear subspace spanned by the columns of the reachability matrix.
By introducing coordinates that span that space the equations for a linear
system can be written as
d xc A11 A12 xc B1
= + u
dt xc̄ 0 A22 xc̄ 0
where the states x o are observable and x ō not observable (quiet) Combin-
ing the representations we find that a linear system can be transformed
to the form
A11 0 A13 0 B1
A B
dx 21 A22 A23 A24 2
= x+ u
dt 0 0 A33 0 0
0 0 A43 A44 0
y = ( C1 0 C2 0) x
154
3.7 Linear Time-Invariant Systems
u y
Soc Σ
-
Soc Soc-
--
Soc
y(t) = u(t) + ce t
Y (s) s−1
= =1
U (s) s−1
155
Chapter 3. Dynamics
y(t) = u(t)
dx
=x
dt
3.8 Summary
This chapter has summarized some properties of dynamical systems that
are useful for control. Both input-output descriptions and state descrip-
tions are given. Much of the terminology that is useful for control has also
been introduced.
156
4
Simple Control Systems
4.1 Introduction
Simple examples of analysis and design of control systems are presented
in this chapter. A cruise controller for a car is designed in Section 4.2.
Only differential equations are used and the necessary mathematics is
summarized in Section 3.3. The analysis of more complicated systems can
be simplified significantly by introducing Laplace transforms and trans-
fer functions as described in Section 3.4. The Laplace transforms make it
easy to manipulate the system formally and to derive relations between
different signals. This is used in Section 4.3, which gives a systematic
way of designing PI controllers for first order systems. The key idea for
design is to develop a process model, to introduce a controller with suf-
ficient degrees of freedom. The controller parameters are then chosen to
give a closed loop system with a specified characteristic polynomial. This
approach to design is called pole placement because it gives a closed loop
system with specified closed loop poles. Some basic properties of second
order systems are also given in Section 4.3. Section 4.4 deals with design
of PI and PID controllers for second order systems Section 4.5 treats the
pole placement problem for systems of arbitrary order. This section which
requires more mathematical maturity can be omitted in a first reading.
For the interested reader it gives, however, important insight into the
design problem and the structure of stabilizing controllers. Section 4.5
summarizes the chapter. In most cases feedback is accomplished by us-
ing sensors and actuators. There are cases where feedback is obtained
through clever process design. On example is an ordinary bicycle which
is analyzed in Section 4.6. It is shown that many of the nice properties of
the bicycle are due to a purely mechanical feedback which has emerged
as a result of trial and error over a long period of time.
157
Chapter 4. Simple Control Systems
Modeling
We will model the system by a momentum balance. The major part of the
momentum is the product of the velocity v and the mass m of the car. There
are also momenta stored in the engine, in terms of the rotation of the crank
shaft and the velocities of the cylinders, but these are much smaller than
mv. Let θ denote the slope of the road, the momentum balance can be
written as
dv
m + cv = F − mgθ (4.1)
dt
where the term cv describes the momentum loss due to air resistance and
rolling and F is the force generated by the engine. The retarding force
158
PSfrag replacements
4.2 Cruise Control
Slope of road
due to the slope of the road should similarly be proportional to the sine
of the angle but we have approximated sin θ θ . The consequence of the
approximations will be discussed later. It is also assumed that the force F
developed by the engine is proportional to the signal u sent to the throttle.
Introducing parameters for a particular car, an Audi in fourth gear, the
model becomes
dv
+ 0.02v = u − 10θ (4.2)
dt
where the control signal is normalized to be in the interval 0 ≤ u ≤ 1,
where u = 1 corresponds to full throttle. The model implies that with full
throttle in fourth gear the car cannot climb a road that is steeper than
10%, and that the maximum speed in 4th gear on a horizontal road is
v = 1/0.02 = 50 m/s (180 km/h).
Since it is desirable that the controller should be able to maintain
constant speed during stationary conditions it is natural to choose a con-
troller with integral action. A PI controller is a reasonable choice. Such a
controller can be described by
Z t
u = k(vr − v) + ki (vr − v(τ ))dτ (4.3)
0
dv de d2 v d2 e
=− , = −
dt dt dt2 dt2
It is convenient to differentiate (4.3) to avoid dealing both with integrals
and derivatives. Differentiating the process model (4.2) the term du/dt
159
Chapter 4. Simple Control Systems
can be eliminated and we find the following equation that describes the
closed loop system
d2 e de dθ
+ (0.02 + k) + ki e = 10 (4.4)
dt2 dt dt
We can first observe that if θ and e are constant the error is zero. This
is no surprise since the controller has integral action, see the discussion
about the integral action Section 2.2.
To understand the effects of the controller parameters k and ki we
can make an analogy between (4.4) and the differential equation for a
mass-spring-damper system
d2 x dx
M +D + Kx = 0
dt2 dt
s2 + 2ζ ω 0 s + ω 20 (4.6)
k = 2ζ ω 0 − 0.02
(4.7)
ki = ω 20
160
4.2 Cruise Control
2.5
Velocity error
2
1.5
0.5
0
0 10 20 30 40 50 60 70 80 90 100
Time
0.5
0.4
Control signal
0.3
0.2
PSfrag replacements
0.1
Time
Velocity error 0
0 10 20 30 40 50 60 70 80 90 100
Control signal Time
Figure 4.3 Simulation of a car with cruise control for a step change in the slope of
the road. The controllers are designed with relative damping ζ = 1 and ω 0 = 0.05
(dotted), ω 0 = 0.1 (full) and ω 0 = 0.2 (dashed).
for a simulation where the slope of the road suddenly changes by 4%. No-
tice that the largest velocity error decreases with increasing ω 0 , but also
that the control signal increases more rapidly. In the simple model (4.1)
it was assumed that the force responded instantaneously to the throttle.
For rapid changes there may be additional dynamics that has to be ac-
counted for. There are also physical limitations to the rate of change of the
force. These limitations, which are not accounted for in the simple model
(4.1), limit the admissible value of ω 0 . Figure 4.3 shows the velocity error
and the control signal for a few values of ω 0 . The largest velocity error de-
creases with increasing values of ω 0 . The smaller error is obtained caused
by a more rapid change in the control signal. A reasonable choice of ω 0 is
in the range of 0.1 to 0.2. The performance of the cruise control system
can be evaluated by comparing the behaviors of cars with and without
cruise control. This is done in Figure 4.4 which shows the velocity error
when the slope of the road is suddenly increased by 4%. Notice the drastic
difference between the open and closed loop systems.
With the chosen parameters ω 0 = 0.2 and ζ = 1 we have 2ζ ω 0 = 0.2
and it follows from (4.7) that the parameter c = 0.02 has little influence
on the behavior of the closed loop system since it is an order of mag-
161
Chapter 4. Simple Control Systems
20
Velocity error
15
10
5
PSfrag replacements
0
0 10 20 30 40 50 60 70 80 90 100
Time
Figure 4.4 Simulation of a car with (solid line) and without cruise control (dashed
line) for a step change of 4% in the slope of the road. The controller is designed for
ω 0 = 0.1 and ζ = 1.
162
4.3 Control of First Order Systems
PSfrag replacements
6
r
5
u
d 4
y
3
C ( s)
P( s) 2
Σ 1
−1 0
e 0 2 4 6 8 10 12
Figure 4.5 Result of a step test experiment. The curve is the response to a unit
step. The gain K is the steady state output and the time constant T is the time
required to reach 63% of the steady state value.
163
Chapter 4. Simple Control Systems
d
PSfrag replacements
r e u y
Σ C ( s) Σ P( s)
−1
Since y(∞) = K we find that the gain K can be determined from the
steady state value of the output. It also follows from the above equation
that
y(T ) = (1 − e−1) K = 0.63K
The time constant T is thus the time required for the step response to
reach 63% of its steady state value. Having obtained the gain K and
the time constant T the parameters a and b are given by a = 1/T and
b = K /T.
By repeating the step test for different amplitudes of the input we can
also obtain an assessment of the validity of the linear model. We can thus
conclude that it is relatively easy to obtain a first order model (4.8) by a
step test and also to verify that the model is reasonable.
PI Control
We will now discuss control of a system with the transfer function (4.8).
A block diagram of the closed loop system obtained with error feedback is
shown in Figure 4.6. A PI controller is well suited for a system described
by (4.8). The PI controller is simple, since it has integral action there will
be no steady state error and with proper choices of parameters the closed
loop system will also be relatively insensitive to the model parameters.
the PI controller has the transfer function
ki
C(s) = k + (4.9)
s
kbs + ki b n L (s)
L(s) = P(s) C(s) = = (4.10)
s(s + a) dL (s)
164
4.3 Control of First Order Systems
The closed loop system is of second order and its characteristic polynomial
is
dL (s) + n L (s) = s2 + (a + bk)s + bki . (4.11)
The poles of the closed loop system can be given arbitrary values by choos-
ing the parameters k and ki properly. Intuition about the effects of the
parameters can be obtained from the mass-spring-damper analogy as was
done in Section 4.2 and we find that integral gain ki corresponds to stiff-
ness and that proportional gain k corresponds to damping.
It is convenient to re-parameterize the problem so that the character-
istic polynomial becomes
s2 + 2ζ ω 0 s + ω 20 (4.12)
2ζ ω 0 − a 2ζ ω 0 T − 1
k= =
b K
(4.13)
ω 20 ω 20 T
ki = =
b K
Since the design method is based on choosing the poles of the closed
loop system it is called pole placement. Instead of choosing the controller
parameters k and ki we now select ζ and ω 0 . These parameters have a
good physical interpretation. The parameter ω 0 determines the speed of
response and ζ determines the shape of the response.
If the parameters ω 0 and ζ are known the controller parameters are
given by (4.13). We will now discuss how to choose these parameters to
obtain desired properties of the system. To do this we must develop an
understanding of the responses of systems of second order.
ω 20
G (s) = . (4.14)
s2 + 2ζ ω 0 s + ω 20
165
Chapter 4. Simple Control Systems
1.5
h
0.5
PSfrag replacements
0
0 5 10 15
ω 0t
Figure 4.7 Step responses h for the system (4.14) with the transfer function
ω 20
G ( s) = for ζ = 0 (dotted), 0.1, 0.2, 0.5, 0.707 (dash dotted), 1, 2, 5 and
s2 +2ζ ω 0 s+ω 20
10 (dashed).
ω0 s + βω 0
G (s) = (4.17)
β s2 + 2ζ ω 0 s + ω 20
Notice that the transfer function has been parameterized so that the
steady state gain G (0) is one. Step responses for this transfer function for
166
4.3 Control of First Order Systems
2.5
2
h
1.5
0.5
PSfrag replacements
0
0 1 2 3 4 5 6 7 8 9 10
1.5
0.5
0
h
−0.5
−1
PSfrag replacements
−1.5
−2
0 1 2 3 4 5 6 7 8 9 10
ω 0t
Figure 4.8 Step responses h for the system (4.14) with the transfer function
ω 0 (s+βω 0 )
G ( s) = 2 2 for ω 0 = 1 and ζ = 0.5. The values for β = 0.25 (dotted), 0.5
β (s +2ζ ω 0 s+ω 0 )
1, 2, 5 and 10 (dashed), are shown in the upper plot and β = −0.25, -0.5 -1, -2, -5
and -10 (dashed) in the lower plot.
different values of β are shown in Figure 4.8. The figure shows that the
zero introduces overshoot for positive β and an undershoot for negative
β . Notice that the effect of β is most pronounced if β is small. The effect
of the zero is small if hβ h > 5. The behavior shown in the figures can be
understood analytically. The transfer function G (s) given by (4.17) can be
written as
ω0 s + βω 0 ω 20 1 sω 0
G (s) = = +
β s + 2ζ ω 0 s + ω 0
2 2 s + 2ζ ω 0 s + ω 0
2 2 β s + 2ζ ω 0 s + ω 20
2
(4.18)
Let h0(t) be the step response of the transfer function
ω 20
G0 (s) =
s2 + 2ζ ω 0 s + ω 20
It follows from (4.18) that the step response of G (s) is
1 dh0(t)
h(t) = h0(t) + (4.19)
βω 0 dt
167
Chapter 4. Simple Control Systems
It follows from this equation that all step responses for different values of
β go through the point where dh0/dt is zero. The overshoot will increase
for positive β and decrease for negative β . It also follows that the effect
of the zero is small if hβ h is large. The largest magnitude of dh/dt is ap-
proximately 0.4ω 0/2.7, which implies that the largest value of the second
term is approximately 0.4/β . The term is thus less than 8% if hβ h is larger
than 5.
Notice in Figure 4.8 that the step response goes in the wrong direction
initially when β is negative. This phenomena is called inverse response,
can also be seen from (4.19). When β is negative the transfer function
(4.17) has a zero in the right half plane. Such are difficult to control
and they are called non-minimum phase system, see Section 3.5. Several
physical systems have this property, for example level dynamics in steam
generators (Example 3.21), hydro-electric power stations (Example 3.20),
pitch dynamics of an aircraft (Example 3.22) and vehicles with rear wheel
steering.
168
4.3 Control of First Order Systems
1.4
1.2
0.8
y
0.6
0.4
0.2
PSfrag replacements
0
0 1 2 3 4 5 6 7 8 9 10
t/ T
Figure 4.9 Step responses of the closed loop system with ζ = 0.5 and ω 0 T = 1
(dashed), 2, 5 and 10 (dash-dotted). The step response of the open loop system is
shown in dashed lines.
to critical damping, compare with Figure 4.7. The overshoot will be ex-
aggerated because of the zero in the numerator of the transfer function
(4.20). Compare with Figure 4.8 which shows the effect of the zero on the
step respoonse. The zero vanishes if 2ζ ω 0 = a as ω 0 increases towards
infinity the zero moves from −∞ to −ω 0 /2ζ . A comparison of the transfer
function (4.20) with the
ω0 ω 0T
β= =
2ζ ω 0 − a 2ζ ω 0 T − 1
169
Chapter 4. Simple Control Systems
The reason for the overshoot is that the controller reacts quite violently
on a step change in the reference. By changing the controller to
Z t
u(t) = − ky(t) + ki (r(τ ) − y(τ ))dτ (4.22)
0
ki
U (s) = − kY (s) + ( R(s) − Y (s)) (4.23)
s
Combining this equation with the equation (4.8) which describes the pro-
cess we find that
bki ω 20
Y (s) = R( s) = R(s) (4.24)
s2 + (a + bk)s + bki s2 + 2ζ ω 0 s + ω 20
P(s) s
G yd (s) = = 2
1 + P(s) C(s) s + (a + bk)s + bki
bs b ω 0s
= 2 =
s + 2ζ ω 0 s + ω 0
2 ω 0 s + 2ζ ω 0 s + ω 20
2
We will first consider the effect of parameter ω 0 . Figure 4.11 shows the
gain curves of the Bode diagram for different values of ω 0 . The figure
shows that disturbances of high and low frequencies are reduced sig-
nificantly and that the disturbance reduction is smallest for frequencies
170
PSfrag replacements
PSfrag replacements
C ( s)
P( s)
Σ 4.3 Control of First Order Systems
r
e r e u y
u Σ C ( s) P( s)
y
−1
d
−1
d
r e u y
ki
Σ s
Σ P
−1
d
0
10
−2
10
−4
10
−3 −2 −1 0 1 2 3
10 10 10 10 10 10 10
Figure 4.11 Gain curves for the transfer function from load disturbance to process
output for b = 1, ζ = 1 and ω 0 = 0.2 dotted, ω 0 = 1.0, dashed and ω 0 = 5 full.
around ω 0 , they may actually be amplified. The figure also shows that
the disturbance rejection at low frequencies is drastically influenced by
the parameter ω 0 but that the reduction of high frequency disturbances
is virtually independent of ω 0 . It is easy to make analytical estimates
because we have
bs s
G yd (s) 2 =
ω0 bki
for small s, where the second equality follows from (4.13). It follows from
this equation that it is highly desirable to have a large value of ω 0 . A
large value of ω 0 means that the control signal has to change rapidly. The
largest permissible value of ω 0 is typically determined by how quickly the
control signal can be changed, dynamics that was neglected in the simple
171
Chapter 4. Simple Control Systems
model (4.8) and possible saturations. The integrated error for a unit step
disturbance in the load disturbance is
Z ∞
1 b 1
IE = e(t)dt = lim E(s) = lim G yd = 2 =
0 s→0 s→0 s ω0 bki
b
max h G yd (iω )h = h G yd (iω 0 )h =
2ζ ω 0
The closed loop system obtained with PI control of a first order system
is of second order. Before proceeding we will investigate the behavior of
second order systems.
172
4.4 Control of Second Order Systems
b1 s + b2
P(s) = (4.25)
s2 + a 1 s + a 2
Pole placement design will now be applied to derive PD, PI and PID
controllers for this process.
PD control
We will start by investigating PD control of the process (4.25). A PD
controller with error feedback has the transfer function
C(s) = k + kd s
(b1 s + b2 )( kd s + k) n L (s)
L(s) = P(s) C(s) = =
s2 + a 1 s + a 2 dL (s)
Requiring that this polynomial has the same roots as the standard second
order polynomial
s2 + 2ζ ω 0 s + ω 20
we get the following equations
a1 + b1 k + b2 kd = (1 + b1 k − d)2ζ ω 0
a2 + b2 k = (1 + b1 k − d)ω 20
173
Chapter 4. Simple Control Systems
PC (b1 s + b2 )( kd s + k)
G yr (s) = = 2
1 + PC s + 2ζ ω 0 s + ω 20
Notice that there will be a steady state error unless a2 = 0. The steady
state error is small if ha2 h << ω 20 . Also notice that the numerator has
two zeros, one is equal to the process zero s = −b2 /b1 the other is due to
the controller. The zero caused by the controller can be eliminated if the
controller based on error feedback can be replaced with the controller
which has two degrees of freedom. The transfer function from reference
to output for the closed loop system then becomes
k(b1 s + b2 )
G yr (s) =
s2 + 2ζ ω 0 s + ω 20
Notice that this transfer function does not have a zero. Block diagrams
for the system with error feedback and with two degrees of freedom are
shown in Figure 4.12.
ki ks + ki
C(s) = k + =
s s
( ks + ki )(b1 s + b2 ) n L (s)
L(s) = P(s) C(s) = =
s3 + a 1 s2 + a 2 s dL (s)
174
PSfrag replacements
PSfrag replacements
k + kd s
P
Σ 4.4 Control of Second Order Systems
r
e r e u y
u Σ k + kd s P
y
−1
d −1
d
r e u y
Σ k Σ P
kd s
−1
d
Figure 4.12 Block diagrams of system with PD control based on error feedback
(above) and with a PD controller with two degrees of freedom (below). Compare
with Figure 4.10.
a1 + b1 k = α 0 + 2ζ 0ω 0
a2 + b1 ki + b2 k = ω 20 + 2α 0ζ 0ω 0
b2 ki = α 0ω 20
Since there are three equations and only two unknowns the problem can-
not be solved in general. To have a solution we can let α 0 or ω 0 be free
parameters. If b1 = 0 and b2 = 0 the equation has the solution
a1
ω0 =
α 0 + 2ζ 0
(1 + 2α 0ζ 0 )ω 20 − a2
k= (4.29)
b2
α 0ω 30
ki =
b2
175
Chapter 4. Simple Control Systems
(α 0 + 2ζ 0)ω 0 − a1
k=
b1
α 0ω 0
3
ki =
b2
In both cases we find that with PI control of a second order system there
is only one choice of ω 0 that is possible. The performance of the closed
loop system is thus severely restricted when a PI controller is used.
PID Control
Assume that the process is characterized by the second-order model
b1 s + b2
P(s) = (4.30)
s2 + a1 s + a 2
This model has four parameters. It has two poles that may be real or com-
plex, and it has one zero. This model captures many processes, oscillatory
systems, and systems with right half-plane zeros. The right half-plane
zero can also be used as an approximation of a time delay. Let controller
be
ki
U (s) = k(bR(s) − Y (s)) + ( R(s) − Y (s)) + kd s(cR(s) − Y (s))
s
( kd s2 + ks + ki )(b1 s + b2 ) n L (s)
L(s) = =
s(s2 + a1 s + a2 ) dL (s)
176
4.4 Control of Second Order Systems
(s + αω 0 )(s2 + 2ζ 0ω 0 s + ω 20 )
a1 + b2 kd + b1 k
= (α + 2ζ 0)ω 0
1 + b1 kd
a2 + b2 k + b1 ki
= (1 + 2αζ 0 )ω 20
1 + b1 kd
b2 ki
= αω 30
1 + b1 kd
b1 s2 + b2 s
G yd =
(s + αω 0 )(s2 + 2ζ 0ω 0 s + ω 20 )
177
Chapter 4. Simple Control Systems
1− s
P(s) =
s2 + 1
This system has one right half-plane zero and two undamped complex
poles. The process is difficult to control.
s3 + 2s2 + 2s + 1.
kv
P(s) = e−sTd (4.32)
s(1 + sT )
where the time delay Td is much smaller than the time constant T. Since
the time constant T is small it can be neglected and the design can be
based on the second order model
kv
P(s) (4.33)
s(1 + sT )
A PI controller for this system can be obtained from Equation (4.29) and
we find that a closed loop system with the characteristic polynomial (4.28)
can be obtained by choosing the parameter ω 0 equal to 1/(α + 2ζ 0)T. Since
Td << T it follows that ω 0 Td << 1 and it is reasonable to neglect the
time delay.
If the approximation (4.33) it is possible to find a PID controller that
gives the closed loop characteristic polynomial with arbitrarily large val-
ues of ω 0 . Since the real system is described by (4.32) the parameter ω 0
must be chosen so that the approximation (4.33) is valid. This requires
that the product ω 0 Td is not too large. It can be demonstrated that the
approximation is reasonable if ω 0 Td is smaller than 0.2.
Summarizing we find that it is possible to obtain the characteristic
polynomial (4.28) with both PI and PID control. With PI control the pa-
rameter ω 0 must be chosen as 1/(α + 2ζ 0 )T. With PID control the pa-
rameter instead can be chosen so that the product ω 0 Td < 1 is small, e.g.
178
4.5 Control of Systems of High Order*
The method for control design used in the previous sections can be charac-
terized in the following way. Choose a controller of given complexity, PD,
PI or PID and determine the controller parameters so that the closed loop
characteristic polynomial is equal to a specified polynomial. This tech-
nique is called pole placement because the design is focused on achieving
a closed loop system with specified poles. The zeros of the transfer func-
tion from reference to output can to some extent be influenced by choos-
ing a controller with two degrees of freedom. We also observed that the
complexity of the controller reflected the complexity of the process. A PI
controller is sufficient for a first order system but a PID controller was
required for a second order system. Choosing a controller of too low order
imposed restrictions on the achievable closed loop poles. In this section
we will generalize the results to systems of arbitrary order. This section
also requires more mathematical preparation than the rest of the book.
Consider a system given by the block diagram in Figure 4.6. Let the
process have the transfer function
where a(s) and b(s) are polynomials. A general controller can be described
by
f (s) U (s) = −g(s) Y (s) + h(s) R(s) (4.35)
where f (s), g(s) and h(s) are polynomials. The controller given by (4.35) is
a general controller with two degrees of freedom. The transfer function
from measurement signal y to control signal u is −g(s)/ f (s) and the
transfer function from reference signal r to control signal u is h(s)/ f (s).
For a system with error feedback we have g(s) = h(s). Elimination of
U (s) between Equations (4.34) and (4.35) gives
a(s) f (s) + b(s)g(s) Y (s) = b(s)h(s) R(s) + b(s) f (s) D (s) (4.36)
179
Chapter 4. Simple Control Systems
Notice that this only depends on the polynomials f (s) and g(s). The design
problem can be stated as follows: Given the polynomials a(s), b(s) and c(s)
find the polynomials f (s) and g(s) which satisfies (4.37). This is a well
known mathematical problem. It will be shown in the next section that
the equation always has a solution if the polynomials a(s) and b(s) do not
have any common factors. If one solution exists there are also infinitely
many solutions. This is useful because it makes it possible to introduce
additional constraints. We may for example require that the controller
should have integral action.
A Naive Solution
To obtain the solution to the design problem the equation (4.37) must be
solved. A simple direct way of doing this is to introduce polynomials f
and g with arbitrary coefficients, writing equating coefficients of equal
powers of s, and solving the equations. This procedure is illustrated by
an example.
f =1
g = s3 + (2a − 1)s2 + (2a2 − 2)s + a3 − 1
f = s + 2a − 2
g = (2a2 − 4a + 3)s + a3 − 2a + 2
180
4.5 Control of Systems of High Order*
3x + 2 y = 1,
181
Chapter 4. Simple Control Systems
ax + by = c (4.38)
has a solution if and only if the greatest common factor of a and b divides
c. If the equation has a solution x0 and y0 then x = x0 − bn and y = y0 + an,
where n is an arbitrary integer, is also a solution.
PROOF 4.1
We will first determine the largest common divisor of the polynomials a
and b by a recursive procedure. Assume that the degree of a is greater
than or equal to the degree of b. Let a0 = a and b0 = b. Iterate the
equations
an+1 = bn
bn+1 = an mod bn
ax + by = bn
An Algorithm
The following is a convenient way of organizing the recursive computa-
tions. With this method we also obtain the minimum degree solution to
the homogeneous equation.
ax + by = 1
(4.39)
au + bv = 0
182
4.5 Control of Systems of High Order*
where g is the greatest common divisor of a and b and u and v are the
minimal degree solutions to the homogeneous equation These equations
can be written as
x ya 1 0 1 x y
=
u v b 0 1 0 u v
g0 + qa
C=
f 0 − qb
PROOF 4.2
The loop transfer function obtained with the controller C is
b(g0 + qa)
L = PC =
a( f 0 − qb)
183
Chapter 4. Simple Control Systems
we have
a f + bg = c
Youla-Kučera Parameterization
Theorem 4.2 characterizes all controllers that give a closed loop system
with a given characteristic polynomial. We will now derive a related re-
sult that characterizes all stabilizing controllers. To start with we will
introduce another representation of a transfer function.
Stable rational functions are also a ring. This means that Theorem 4.1
also holds for rational functions. A fractional representation of a transfer
function P is
B
P=
A
where A and B are stable rational transfer functions. We have the fol-
lowing result.
184
4.5 Control of Systems of High Order*
v
PSfrag replacements
Q B Σ −A
1
− G0 Σ P
F0
G0 + QA
C= (4.40)
F0 − QB
PROOF 4.3
The loop transfer function obtained with the controller C is
B ( G0 + QA)
L = PC =
A( F0 − QB )
we have
A( F0 − QB ) + B ( G0 + QA) AF0 + B G0
1+ L = =
A( F0 − QB ) A( F0 − QB )
Since the rational function AF0 + B G0 has all its zeros in the left half
plane the closed loop system is stable. Let C = G / F be any controller
that stabilizes the closed loop system it follows that
AF + B G = C
is a stable rational function with all its zeros in the left half plane. Hence
A B
F+ G=1
C C
185
Chapter 4. Simple Control Systems
B
F = F0 − Q = F0 − B Q̄
C
A
G = G0 − Q = G0 − A Q̄
C
where Q is a stable rational function because C has all its zeros in the
left half plane.
It follows from Equation (4.40) that the control law can be written as
U G G0 + QA
=− =−
Y F F0 − QB
or
F0 U = − G0 Y + Q( B U − AY )
The Youla-Kučera parameterization theorem can then be illustrated by
the block diagram in Figure 4.13. Notice that the signal v is zero. It
therefore seems intuitively reasonable that a feedback based on this signal
cannot make the system unstable.
Modeling
When modeling a physical object it is important to keep the purpose of
the model in mind. In this modeling exercise we will derive simple models
that can be used to discuss the basic balancing and steering problmes for
186
4.6 Bicycle Dynamics and Control
Figure 4.14 Schematic picture of a bicycle. There are two coordinate systems the
inertial frame x yz and the system ξηζ that is fixed to the bicycle. The coordinates
of the rear wheel are x and y. The orientation of the bicycle is given by yaw angle
ψ , roll angle φ and steer angle δ .
187
Chapter 4. Simple Control Systems
Figure 4.15 Bicycle kinematics. The wheel base is b and a is the ξ coordinate of
the center of mass.
dγ r dγ f
= sin δ f .
dt dt
The system has three degrees of freedom beacause there are 8 configura-
tion variables and 5 nonholonomic constraints. This reduces to two if the
rider is assumed to be rigidly attached to the frame.
A Basic Model
We will first make additional simplifying assumptions. It is assumed that
the rider has fixed position and orientation realtive the frame. The steer
angle δ is assumed to be the control variable. The rotational degree of
freedom associated with the front fork then disappears and the system
only has one degree of freedom, which can be chosen as the roll angle ϕ .
The roll of the front fork will be neglected which means that δ f = δ and
all angles are assumed to be so small that the equations can be linearized.
The forward velocity at the rear wheel V is assumed to be constant.
To describe the equations of motion we assume that the bicycle rolls
on the horisontal x y plane. Introduce a coordinate system that is fixed
to the bicycle with the ξ -axis through the contact points of the wheels
with the ground, the η -axis horizontal and the ζ -axis vertical as shown
in Figure 4.14. The coordinate system is not fixed in inertial space. An
observer fixed to the coordinate system will therefore experience forces
due to the motion of the coordinate system. The motion of the coordinate
system can be determined from kinematics, see Figure 4.15. We find that
188
4.6 Bicycle Dynamics and Control
the coordinate system rotates around the point O with the angular velocity
ω = V δ /b.
Let m be the mass of the system, b the wheel base, J the moment of
inertia with respect to the ξ -axis, D the product of inertia with respect
to the ξζ plane. Furthermore let the ξ and ζ coordinates of the center of
mass be a and h. The angular momentum of the system is
dϕ dϕ V D dδ
L=J − Dω = J −
dt dt b dt
The torques acting on the system are due to gravity and centrifugal action.
The conservation of angular momentum can be written as
d2ϕ DV dδ mV 2 h
J − = mg hϕ + δ (4.41)
dt2 b dt b
where the first term of the right hand side is the torque generated by
gravity and the second term is the torque generated by the centrifugal
forces. It follows from (4.41) that the transfer function from steer angle
δ to roll angle ϕ is is
mV h
V ( Ds + mV h) VD s+ D
Gϕδ (s) = = (4.42)
b( Js2 − mg h) bJ 2 mg h
s −
J
189
Chapter 4. Simple Control Systems
Figure 4.16 Schematic picture of the front fork. The angle λ is called the caster
angle.
contact point and the projection of the axis of rotation of the front fork
assembly c is called the the trail. The distance c is called the trail. The
properties of the bicycle are strongly influenced by the value of c.
The main forces acting on the front fork are the contact forces. With
a positive trail the normal force turns the wheel towards the lean when
the bicycle is tilted. The friction forces acts in the opposite direction. The
steering properties of a bicycle depend critically on the trail. A large trail
increases stability but make the steering less agile.
Let T be the torque applied on the front wheel assembly by the driver.
A simple linearized model is.
δ = k1 T − k2ϕ . (4.43)
190
4.6 Bicycle Dynamics and Control
T
PSfrag replacements
T Front fork
k1
δ ϕ
k2 Σ Frame
−1
Figure 4.17 Block diagram of a bicycle with a front fork, T is handle bar torque,
ϕ roll angle and δ steer angle.
d2ϕ DV k2 dϕ mV 2 hk2 DV k1 dT mV 2 k1 h
J 2
+ + − mg h ϕ = + T (4.44)
dt b dt b b dt b
DV k2 mV 2 hk
2
Js2 + s+ − mg h = 0
b b
We can thus conclude that because of the feedback created by the design of
the front fork the bicycle will be stable, even in the absence of rider applied
steer torque, provided that the velocity is sufficiently large. The velocity V c
is called the critical velocity. Reasonable parameters for a normal bicycle
with non-articulating rider are m = 80 kg, a = 0.3 m, b = 1 m, h = 1.2
m, J = mh2 kgm2 , k1 = 0.03 N−1m−1 and k2 = 4. This gives a critical
velocity of 1.6 m/s.
Biking in a Circle
Useful information about bicycle dynamics can be obtained by driving
it with constant speed V in a circle with radius r0 . To determine the
numerical values of the essential parameters a torque wrench can be used
to measure the torque the driver exerts on the handle bar. In steady state
191
Chapter 4. Simple Control Systems
bg V 2 b V2
T0 = − 1 ϕ 0 = 1 − (4.46)
k1 V 2 Vc2 k1 r0 Vc2
This gives the nonintuitive result that the torque is positive when the
speed is less than the critical speed and negative when the speed is larger
than the critical speed. The steady state analysis implies that no torque is
required if the bicycle is driven at the critical velocity. This non-intuitive
result can be used to determine the critical speed. Notice that the calcula-
tion is based on the assumption that the rider is lined up with the bicycle,
small changes in the rider orientation will change the torque significantly.
Gyroscopic Effects
The gyroscopic action of the rotation of the front wheel can be taken into
account in the model of the front fork. The model (4.43) is then replaced
by
dϕ
δ = k1 T − k2ϕ − k3
dt
and the equation for the closed loop system becomes
Analysis of the simple model given by Equations (4.41) and (4.43) shows
that the damping is improved but the condition for stability V > V c does
not change.
Rear-wheel Steering
The analysis performed shows that feedback analysis gives substantial
insight into behavior of bicycles. Feedback analysis can also indicate that
a proposed system may have substantial disadvantages that are not ap-
parent from static analysis. It is therefore essential to consider feedback
and dynamics at an early stage of design. These issues can be illustrated
very nicely by investigating a bicycle with rear wheen steering. There are
advantages in having rear-wheel steering on recumbent bicycles because
the design of the drive is simpler. Again we quote from Whitt and Wilson
Bicycling Science:
192
4.6 Bicycle Dynamics and Control
Many people have seen theoretical advantages in the fact that front-
drive, rear-steered recumbent bicycles would have simpler transmis-
sions than rear-driven recumbents and could have the center of mass
nearer the front wheel than the rear. The U.S. Department of Trans-
portation 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.
The reason for this is that a bicycle with rear-wheel steering has dynamics
which makes it very difficult to ride because the transfer function from
steering angle to tilt given by Equation (??) is replaced by
amV h −s + Va
Gϕδ (s) = , (4.47)
bJ s2 − mJg h
One way to derive this is to simply reverse the sign of the velocity in
Equation (??). The transfer function (4.47) which has both poles and zeros
in the right half plane and is very difficult to control. This can be seen
by investigating the closed loop system obtained when (4.43) is combined
with the front fork model (4.43). The closed loop system then becomes
Manoevring
Having obtained some insight into stabilization of bicycles we will now
turn to the problem of steering. A central problem is to determine how
the path of the bicycle is influenced by the handle bar torque. Later we
will also investigate the effects of leaning.
193
Chapter 4. Simple Control Systems
The first step is to investigate how the steering torque influences the
steering angle. It follows from the block diagram in Figure 4.17 that
k2
Gδ T (s) =
1 + k1 Gϕδ (s)
where Gϕδ (s) is the transfer function from steering torque T to tilt ϕ
given by (4.42). We thus find that
mg h
k2 s2 −
Gδ T (s) = J (4.48)
2
amhk1 V0 mg h V02
s + s+ − 1
bJ J Vc2
Notice that the poles of the transfer function Gϕδ appear as zeros of the
transfer function Gδ T . The need for stabilization of the bicycle thus implies
that the steering dynamics has a zero in the right half plane. This imposes
limitations in the maneuvrability of the bicycle.
To determine how the path is influenced by the steering angle we
introduce the coordinate system in Figure 4.14 where x and y are the
coordinates of the rear wheel contact point. The equations are then lin-
earized by assuming small deviations from a straight line path. It then
follows
dy
= Vψ
dt (4.49)
dψ V
= δ
dt b
The transfer function from steering angle δ to path deviation y is
V2
G yδ (s) = (4.50)
bs2
Combining this with Equation (4.48) gives the following transfer function
from steering torque T to y.
ak2 V s2 − mg h/ J
G yT (s) =
b amhk1 V mg h V 2
s2 s2 + s+ ( 2 − 1)
bJ J Vc
Figure 4.18 shows the path of a bicycle when a constant torque is applied
to the handle bar. The inverse response nature of the path is clearly visible
in the figure. This can be explained physically as follows. When a steering
torque is applied, the bicycle’s contact line with the ground will initially
194
4.6 Bicycle Dynamics and Control
PSfrag replacements
y
δ
ϕ
t
Figure 4.18 Path deviation y, tilt angle ϕ and steering angle δ of a bicycle when
a constant torque is applied to the handle bar. Perhaps better to show a complete
simulation?
turn in the direction of the applied torque. When the bicycle turns there
will be a centripetal force that tilts the bicycle around the x axis. This
tilt generates a torque on the front fork which turns the front fork in the
opposite direction. Once we see the curves it is straight forward to come
up with a physical explanation of what happens. It is however difficult to
apply causal reasoning directly because of the closed loop nature of the
system. The physical argument also indicates a remedy because the tilt
generated by the centripetal force can be opposed by gravitational forces
if the biker bends his upper torso relative to the frame.
The inverse response nature of the response to steering torque has
caused many motorcycle accidents. To explain this we will visualize a mo-
torcycle driven along a straight path. Assume that a car might suddenly
appear from the right. The intuitive reaction is then to try to steer away
from the car. The motor cycle will initially do so when the steering torque
is applied but because of the nature of the steering dynamics the motor
195
Chapter 4. Simple Control Systems
bike will steer into the car as indicated in Figure 4.18. Again the remedy
is to simultaneously lean and counter-steer. Counter-steering consists of
a momentary turn of the handlebar in the opposite direction of the in-
tended travel. This establishes a lean which turns the vehicle away from
the danger( or obstacle that has appeared in front of the rider). Failure to
counter-steer often occurs with novice riders and is responsible for a sig-
nificant number of motor cycling accidents and resulting deaths. Leaning
suffice without counter-steering in many normal traveling conditions, like
routine lane changes. Counter-steer is, however, necessary for fast lane
changes because the dynamics of leaning alone has a long response time.
lean. Since a large motor cycle is heavy a significant lean is required.
Effects of Leaning
Sofar it has been assumed that the rider sits rigidly on the bicycle without
leaning. This is a strong simplification. A simple way of describing the
effect of leaning is to consider the rider as composed of two rigid pieces,
the lower body and the upper body, where the upper body can be rotated
as is illustrated in Figure 4.19. Let the upper body lean the angle be φ
relative to a plane through the bicycle. The linearized momentum balance
given by (??) is then replaced by
d2ϕ d 2φ
J + J 2 = mg hϕ + m2 g h2φ + Fh (4.51)
dt2 dt2
where m2 is the mass of the upper body, J2 is the moment of inertia of the
upper body with respect to the line where the wheels make contact with
the ground, and l2 is the distance from the ground to the center of mass
of the upper body. Combining this equation with the expression (??) for
196
4.6 Bicycle Dynamics and Control
the force and the model of the front fork (4.43) gives the following model
of the bicycle
d2ϕ dϕ dT dφ
2
+ a1 + a2ϕ = b11 + b12 T + b21 + b22φ
dt dt dt dt
δ = − k2ϕ + k1 T
a ak2 ak1 (4.52)
α = δ =− ϕ+ T
b b b
ak2 V ak1 V
Vy = V α = − ϕ+ T
b b
where the coefficients are given by
amV hk2 mg h aV
a1 = =
bJ J Vc2
mV 2 k2 h mg h mg h V 2
a2 = − = −1
bJ J J Vc2
amV hk1 k1
b11 = = a1
bJ k2 (4.53)
mV 2 k1 h mg h k1 V 2 k1 mg h
b12 = = = a 2 +
bJ J k2 Vc2 k2 J
J2
b21 =−
J
m2 g h 2 mg h m2 h2
b22 = =
J J mh
model of a bicycle is a second order system with two inputs, the steering
torque T and the lean φ .
In steady state Equation (4.52) reduces to
Assuming for example that the bicycle moves in a straight path we have
δ = 0. Straight forward calculations give
k2 m2 h2
T=− φ
k1 mh
m2 h 2
ϕ =− φ
mh
A positive tilt thus requires a negative torque on the handle bars and a
negative lean of the bicycle. This is something you can easily try yourself
197
Chapter 4. Simple Control Systems
Summary
There are several interesting general conclusions that can be drawn from
the analysis of this section. The bicycle is a good illustration of the dif-
ficulties in reasoning intuitively about feedback systems. Simple causal
reasoning fails because of the closed loop nature of the system. The fact
that the bicycle is self stabilizing is certainly not a trivial matter. An-
other interesting feature is that much insight can be derived from simple
mathematical models. It is however important to make the proper sim-
plifications. A more subtle observation is that stabilization of the bicycle
introduces difficulties in steering because of the unstable pole appears as
a zero in the steering dynamics. This is actually a general property that
appears in control of many unstable systems. The last observation is that
the difficulty with the steering dynamics can be eliminated by introduc-
ing an extra control variable. This is a manifestation of a very general
result that difficulties in control arising from the appearance of zeros in
the right half plane can be eliminated by introducing additional sensors
or actuators.
4.7 Summary
198
4.7 Summary
199
5
Feedback Fundamentals
5.1 Introduction
Fundamental properties of feedback systems will be investigated in this
Chapter. We begin in Section 5.2 by discussing the basic feedback loop and
typical requirements. This includes the ability to follow reference signals,
effects of load disturbances and measurement noise and the effects of pro-
cess variations. It turns out that these properties can be captured by a
set of six transfer functions, called the Gang of Six. These transfer func-
tions are introduced in Section 5.3. For systems where the feedback is
restricted to operate on the error signal the properties are characterized
by a subset of four transfer functions, called the Gang of Four. Properties
of systems with error feedback and the more general feedback configura-
tion with two degrees of freedom are also discussed in Section 5.3. It is
shown that it is important to consider all transfer functions of the Gang
of Six when evaluating a control system. Another interesting observation
is that for systems with two degrees of freedom the problem of response
to load disturbances can be treated separately. This gives a natural sepa-
ration of the design problem into a design of a feedback and a feedforward
system. The feedback handles process uncertainties and disturbances and
the feedforward gives the desired response to reference signals.
Attenuation of disturbances are discussed in Section 5.4 where it is
demonstrated that process disturbances can be attenuated by feedback
but that feedback also feeds measurement noise into the system. It turns
out that the sensitivity function which belongs to the Gang of Four gives
a nice characterization of disturbance attenuation. The effects of process
variations are discussed in Section 5.5. It is shown that their effects are
well described by the sensitivity function and the complementary sensi-
tivity function. The analysis also gives a good explanation for the fact that
200
5.2 The Basic Feedback Loop
d n
r e u v x y
F Σ C Σ P Σ
−1
Controller Process
201
Chapter 5. Feedback Fundamentals
PSfrag replacementsw z
P
u y
Figure 5.2 An abstract representation of the system in Figure 5.1. The input u
represents the control signal and the input w represents the reference r, the load
disturbance d and the measurement noise n. The output y is the measured variables
and z are internal variables that are of interest.
202
5.2 The Basic Feedback Loop
Disturbances
Attenuation of load disturbances is often a primary goal for control. This is
particularly the case when controlling processes that run in steady state.
Load disturbances are typically dominated by low frequencies. Consider
for example the cruise control system for a car, where the disturbances are
the gravity forces caused by changes of the slope of the road. These distur-
bances vary slowly because the slope changes slowly when you drive along
a road. Step signals or ramp signals are commonly used as prototypes for
load disturbances disturbances.
Measurement noise corrupts the information about the process vari-
able that the sensors delivers. Measurement noise typically has high fre-
quencies. The average value of the noise is typically zero. If this was not
the case the sensor will give very misleading information about the pro-
cess and it would not be possible to control it well. There may also be
dynamics in the sensor. Several sensors are often used. A common situa-
tion is that very accurate values may be obtained with sensors with slow
dynamics and that rapid but less accurate information can be obtained
from other sensors.
Actuation
The process is influenced by actuators which typically are valves, motors,
that are driven electrically, pneumatically, or hydraulically. There are of-
ten local feedback loops and the control signals can also be the reference
variables for these loops. A typical case is a flow loop where a valve is
controlled by measuring the flow. If the feedback loop for controlling the
flow is fast we can consider the set point of this loop which is the flow
as the control variable. In such cases the use of local feedback loops can
thus simplify the system significantly. When the dynamics of the actua-
tors is significant it is convenient to lump them with the dynamics of the
process. There are cases where the dynamics of the actuator dominates
process dynamics.
Design Issues
Many issues have to be considered in analysis and design of control sys-
tems. Basic requirements are
• Stability
• Ability to follow reference signals
• Reduction of effects of load disturbances
• Reduction of effects of measurement noise
• Reduction of effects of model uncertainties
203
Chapter 5. Feedback Fundamentals
P PC PCF
X = D− N+ R
1 + PC 1 + PC 1 + PC
P 1 PCF
Y= D+ N+ R (5.1)
1 + PC 1 + PC 1 + PC
PC C CF
U =− D− N+ R.
1 + PC 1 + PC 1 + PC
PCF PC P
1 + PC 1 + PC 1 + PC
(5.2)
CF C 1
,
1 + PC 1 + PC 1 + PC
204
5.3 The Gang of Six
The transfer functions in the first column give the response of process
variable and control signal to the set point. The second column gives the
same signals in the case of pure error feedback when F = 1. The transfer
function P/(1 + PC) in the third column tells how the process variable
reacts to load disturbances the transfer function C/(1 + PC) gives the
response of the control signal to measurement noise.
Notice that only four transfer functions are required to describe how
the system reacts to load disturbance and the measurement noise and
that two additional transfer functions are required to describe how the
system responds to set point changes.
The special case when F = 1 is called a system with (pure) error
feedback. In this case all control actions are based on feedback from the
error only. In this case the system is completely characterized by four
transfer functions, namely the four rightmost transfer functions in (5.2),
i.e.
PC
, the complementary sensitivity function
1 + PC
P
, the load disturbance sensitivity function
1 + PC
(5.3)
C
, the noise sensitivity function
1 + PC
1
, the sensitivity function
1 + PC
These transfer functions and their equivalent systems are called the
Gang of Four. The transfer functions have many interesting properties
that will be discussed in then following. A good insight into these prop-
erties are essential for understanding feedback systems. The load distur-
bance sensitivity function is sometimes called the input sensitivity func-
tion and the noise sensitivity function is sometimes called the output
sensitivity function.
205
Chapter 5. Feedback Fundamentals
1 1 1
0 0 0
0 10 20 30 0 10 20 30 0 10 20 30
C F /(1 + PC) C/(1 + PC) 1/(1 + PC)
PSfrag replacements 1.5 1.5 1.5
1 1 1
0 0 0
0 10 20 30 0 10 20 30 0 10 20 30
Figure 5.3 Step responses of the Gang of Six for PI control k = 0.775, Ti = 2.05
of the process P( s) = ( s + 1)−4 . The feedforward is designed to give the transfer
function (0.5s + 1)−4 from reference r to output y.
problem.
To describe the system properly it is thus necessary to show the re-
sponse of all six transfer functions. The transfer functions can be repre-
sented in different ways, by their step responses and frequency responses,
see Figures 5.3 and 5.4.
Figures 5.3 and 5.4 give useful insight into the properties of the closed
loop system. The time responses in Figure 5.3 show that the feedforward
gives a substantial improvement of the response speed. The settling time
is substantially shorter, 4 s versus 25 s, and there is no overshoot. This is
also reflected in the frequency responses in Figure 5.4 which shows that
the transfer function with feedforward has higher bandwidth and that it
has no resonance peak.
The transfer functions CF /(1 + PC) and − C/(1 + PC) represent the
signal transmission from reference to control and from measurement noise
to control. The time responses in Figure 5.3 show that the reduction in
response time by feedforward requires a substantial control effort. The
initial value of the control signal is out of scale in Figure 5.3 but the
frequency response in 5.4 shows that the high frequency gain of PCF /(1 +
PC) is 16, which can be compared with the value 0.78 for the transfer
function C/(1 + PC). The fast response thus requires significantly larger
control signals.
There are many other interesting conclusions that can be drawn from
Figures 5.3 and 5.4. Consider for example the response of the output to
206
5.3 The Gang of Six
−1 −1 −1
10 10 10
−1 0 1 −1 0 1 −1 0 1
10 10 10 10 10 10 10 10 10
C F /(1 + PC) C/(1 + PC) 1/(1 + PC)
PSfrag replacements
1
10
1 1
10 10
0
0
10
0
10 10
−1 0 1 −1 0 1 −1 0 1
10 10 10 10 10 10 10 10 10
Figure 5.4 Gain curves of frequency responses of the Gang of Six for PI control
k = 0.775, Ti = 2.05 of the process P( s) = ( s + 1)−4 where the feedforward has been
designed to give the transfer function (0.5s + 1)−4 from reference to output.
A Remark
The fact that 6 relations are required to capture properties of the basic
feedback loop is often neglected in literature. Most papers on control only
show the response of the process variable to set point changes. Such a
curve gives only partial information about the behavior of the system. To
get a more complete representation of the system all six responses should
be given. We illustrate the importance of this by an example.
207
Chapter 5. Feedback Fundamentals
Process variable
2
1.5
0.5
0
0 10 20 30 40 50 60
Control signal
1.5
0.5
−0.5
0 10 20 30 40 50 60
208
5.3 The Gang of Six
Output
1.4
1.2
1
y 0.8
0.6
0.4
0.2
0
0 2 4 6 8 10 12 14 16 18 20
Control
1.2
0.8
0.4
0.2
−0.2
0 2 4 6 8 10 12 14 16 18 20
t
1 (s + 1)(s + 0.02)
Y (s) = R(s), U (s) = R(s)
s2 + s + 1 s2 + s + 1
s 1
Y (s) = D (s), U (s) = − D (s)
(s + 0.02)(s2 + s + 1) s2 +s+1
(s + 0.02)(s2 + s + 1) = 0
where the the pole s = −0.02 corresponds the process pole that is canceled
by the controller zero. The presence of the slow pole s = −0.02 which ap-
pears in the response to load disturbances implies that the output decays
very slowly, at the rate of e−0.02t. The controller will not respond to the
209
Chapter 5. Feedback Fundamentals
Output
1.2
0.8
0.6
y
0.4
0.2
−0.2
0 20 40 60 80 100 120 140 160 180 200
Control
0
−0.2
−0.4
−0.8
−1
−1.2
−1.4
0 20 40 60 80 100 120 140 160 180 200
t
Figure 5.7 Response of output y and control u to a step in the load disturbance.
Notice the very slow decay of the mode e−0.02t. The control signal does not respond
to this mode because the controller has a zero s = −0.02. The dashed curves show
the results when the controller is modified to C ( s) = 1 + 0.2/ s.
signal e−0.02t because the zero s = −0.02 will block the transmission of this
signal. This is clearly seen in Figure 5.7, which shows the response of the
output and the control signals to a step change in the load disturbance.
Notice that it takes about 200 s for the disturbance to settle. This can be
compared with the step response in Figure 5.6 which settles in about 10s.
Having understood what happens it is straight forward to modify the
controller. With the controller
0.2
C(s) = 1 +
s
210
yol
x
r=0
d
n 5.4 Disturbance Attenuation
C
P d n
Σ
F r=0 u x yol
C Σ P Σ
d n
r=0 e u x ycl
Σ C Σ P Σ
−1
Figure 5.8 Open and closed loop systems subject to the same disturbances.
controller. The canceled factors do not appear in the loop transfer function
and the sensitivity functions. The canceled modes are not visible unless
they are excited. The effects are even more drastic than shown in the
example if the canceled modes are unstable. This has been known among
control engineers for a long time and a good design rule that cancellation
of slow or unstable process poles by zeros in the controller give very poor
attenuation of load disturbances. Another view of cancellations is given
in Section 3.7.
211
Chapter 5. Feedback Fundamentals
1
10
h S(iω )h
10
PSfrag replacements 10
−1
−2
10
−2 −1 0 1
10 10 10 10
+ ω
Figure 5.9 Gain curve of the sensitivity function for PI control ( k = 0.8, k i = 0.4)
of process with the transfer function P( s) = ( s + 1)−4 . The sensitivity crossover
frequency is indicated by + and the maximum sensitivity by o.
where S(s) is the sensitivity function, which belongs to the Gang of Four.
We thus obtain the following interesting result
The sensitivity function will thus directly show the effect of feedback on
the output. The disturbance attenuation can be visualized graphically by
the gain curve of the Bode plot of S(s). The lowest frequency where the
sensitivity function has the magnitude 1 is called the sensitivity crossover
frequency and denoted by ω sc. The maximum sensitivity
1
Ms = max h S(iω )h = max (5.7)
ω ω 1 + P(iω ) C(iω )
212
5.4 Disturbance Attenuation
30
20
10
y 0
−10
−30
0 10 20 30 40 50 60 70 80 90 100
t
Figure 5.10 Outputs of process with control (full line) and without control (dashed
line).
1 1
S(s) = = . (5.8)
1 + P(s) C(s) 1 + L(s)
The first term represents the effect of the load disturbance and the second
term the effect of measurement noise. Load disturbance are thus attenu-
ated but measurement noise is injected because of the feedback.
213
Chapter 5. Feedback Fundamentals
−1
1/ Ms ω ms
PSfrag replacements
ω sc
Figure 5.11 Nyquist curve of loop transfer function showing graphical interpre-
tation of maximum sensitivity. The sensitivity crossover frequency ω sc and the fre-
quency ω ms where the sensitivity has its largest value are indicated in the figure.
All points inside the dashed circle have sensitivities greater than 1.
P(s)
Gxd (s) = = c 0 + c 1 s + c 2 s2 + ⋅ ⋅ ⋅ (5.9)
1 + P(s) C(s)
dd(t) d2 d(t)
x(t) = c0 d(t) + c1 + c2 + ⋅⋅⋅
dt dt2
*Random Disturbances
Process disturbances can often be described as stationary stochastic pro-
cesses which can be characterized by their power spectral density φ (ω ).
214
5.5 Process Variations
This physical meaningRis that the energy of the signal in the frequency
ω
band ω 1 ≤ ω ≤ ω 2 is 2 ω 12 φ (ω )dω . If φ ol (ω ) is the power spectral density
of the output in open loop the power spectral density of the closed loop
system is
φ cl (ω ) = h S(iω )h2φ ol (ω )
and the ratio of the variances of the open and closed loop systems is
R∞
σ cl2 h S(iω )h2φ ol (ω )dω
= −∞ R ∞
σ cl
2
φ (ω )dω
−∞ ol
h C ∆ Ph < h1 + Lh
which implies
1 + PC(i
h∆ Ph < (5.10)
C
This condition must be valid for all points on the Nyquist curve, i.e point-
wise for all frequencies. The condition for stability can be written as
∆ P(iω ) 1
< (5.11)
P(iω ) hT (iω )h
215
Chapter 5. Feedback Fundamentals
−1
1+ L
PSfrag replacements C∆ P
Figure 5.12 Nyquist curve of a nominal loop transfer function and its uncertainty
caused by process variations ∆ P.
216
5.5 Process Variations
permitted without making the closed loop system unstable. The fact that
the closed loop system is robust to process variations is one of the reason
why control has been so successful and that control systems for complex
processes can indeed be designed using simple models. This is illustrated
by an example.
1
P(s) =
(s + 1)4
The situation illustrated in the figure is typical for many processes, mod-
erately small uncertainties are only required around the gain crossover
frequencies, but large uncertainties can be permitted at higher and lower
frequencies. A consequence of this is also that a simple model that de-
scribes the process dynamics well around the crossover frequency is suf-
ficient for design. Systems with many resonance peaks are an exception
to this rule because the process transfer function for such systems may
have large gains also for higher frequencies.
217
Chapter 5. Feedback Fundamentals
∆P ∆P
P
PSfrag replacements P Σ
PSfrag replacements
− 1+PCPC
−C
Figure 5.14B is
PC
L(s) = ∆P
1 + PC
Equation 5.11 thus simply implies that the largest loop gain is less than
one. Since both blocks are stable it follows from Nyquists stability theorem
that the closed loop is stable. This result which holds under much more
general assumptions is called the small gain theorem.
218
5.5 Process Variations
Y PC
= =T (5.13)
R 1 + PC
Compare with (5.2). The transfer function T which belongs to the Gang
of Four is called the complementary sensitivity function. Differentiating
(5.13) we get
dT C PC T
= = =S
dP (1 + PC)2 (1 + PC)(1 + PC) P P
Hence
d log T dT P
= =S (5.14)
d log P dP T
This equation is the reason for calling S the sensitivity function. The
relative error in the closed loop transfer function T will thus be small
if the sensitivity is small. This is one of the very useful properties of
feedback. For example this property was exploited by Black at Bell labs
to build the feedback amplifiers that made it possible to use telephones
over large distances.
A small value of the sensitivity function thus means that disturbances
are attenuated and that the effect of process perturbations also are negli-
gible. A plot of the magnitude of the complementary sensitivity function
as in Figure 5.9 is a good way to determine the frequencies where model
precision is essential.
Constraints on Design
Constraints on the maximum sensitivities M s and Mt are important to
ensure that closed loop system is insensitive to process variations. Typical
constraints are that the sensitivities are in the range of 1.1 to 2. This has
implications for design of control systems which are illustrated by an
example.
219
Chapter 5. Feedback Fundamentals
by
b
Y (s) = U (s)
s+ a
ki
U (s) = − kY (s) + ( R(s) − Y (s))
s
where U , Y and R are the Laplace transforms of the process input, output
and the reference signal. The closed loop characteristic polynomial is
s2 + (a + bk)s + bki
s2 + 2ζ ω 0 s + ω 20 (5.15)
2ζ ω 0 − a
k=
b
ω 20
ki =
b
and there are no apparent constraints on the choice of parameters ζ and
ω 0 . Calculating the sensitivity functions we get
s(s + a)
S(s) =
s2 + 2ζ ω 0 s + ω 20
(2ζ ω 0 − a)s + ω 20
T (s) =
s2 + 2ζ ω 0 s + ω 20
Figure 5.15 shows clearly that the sensitivities will be large if the pa-
rameter ω 0 is chosen smaller than a. The equation for controller gain
also gives an indication that small values of ω 0 are not desirable because
proportional gain then becomes negative which means that the feedback
is positive.
We can thus conclude that if a closed loop characteristic polynomial of
the form (5.15) with ζ ≤ 1 is chosen it is necessary to have ω 0 ≥ a/(2ζ ) in
order to have a system with reasonable robustness. The response time of
the closed loop system thus must be sufficiently fast. It is however possible
to obtain closed loop system with slower response time by choosing a
closed loop characteristic polynomial with real roots. Let the characteristic
polynomial be
(s + p1)(s + p2)
220
5.5 Process Variations
2
10
0
10
hS(iω )h
−2
10
PSfrag replacements
−4
10
−2 −1 0 1 2
10 10 10 10 10
ω
1
10
0
10
hT (iω )h
−1
10
PSfrag replacements 10
−2
hS(iω )h 10
−3
−2 −1 0 1 2
10 10 10 10 10
ω
Figure 5.15 Magnitude curve for Bode plots of the sensitivity function S (above)
and the complementary sensitivity function T (below) for ζ = 0.7, a = 1 and ω 0 / a =
0.1 (dashed), 1 (solid) and 10 (dotted).
p1 + p 2 − a
k=
b
p1 p2
ki =
b
s(s + a)
S(s) =
(s + p1)(s + p2)
( p 1 + p 2 ) s + p 1 p2
T (s) =
(s + p1)(s + p2)
To guarantee that the sensitivity function is not too large one of the closed
loop poles should be close to a. The complementary sensitivity function
has a zero at
p1 p2
z1 =
p1 + p 2
221
Chapter 5. Feedback Fundamentals
2.5
Ms , M t 2
1.5
PSfrag replacements
1
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
ζ
Figure 5.16 Maximum sensitivities Ms (full line) and Mt (dashed line) as func-
ω 20 s(s+2ζ ω 0 )
tions of relative damping for T ( s) = s2 +2ζ ω 0 s+ω 20
and S ( s) = s2 +2ζ ω 0 s+ω 20
.
If P1 ≤ p2 it follows that
z1 p1
0.5 < = <1
p1 p1 + p 2
This implies that hT (iω )h<2. We can thus conclude that if it is desired
to have a closed loop system with slower response time one of the closed
loop poles should be chosen close to the process pole a. This is another
example of the fact that it is important to choose the closed loop poles
carefully when using pole placement design.
ω 20
T (s) =
s2 + 2ζ ω 0 s + ω 20
s(s + 2ζ ω 0 )
S(s) = 1 − T (s) =
s2 + 2ζ ω 0 s + ω 20
222
5.6 When are Two Processes Similar?
s p
8ζ 2 + 1 + (4ζ 2 + 1) 8ζ 2 + 1
Ms = p
8ζ 2 + 1 + (4ζ 2 − 1) 8ζ 2 + 1
p
1 + 8ζ 2 + 1
wms = ω0
2
p √
1/(2ζ 1 − ζ 2 ) if ζ ≤ 2/2 (5.16)
Mt = √
1 if ζ > 2/2
p √
ω 0 1 − 2ζ 2 if ζ ≤ 2/2
ω mt = √
0 if ζ > 2/2
The relation between the sensitivities and relative damping are shown in
Figure 5.16. The values ζ = 0.3, 0.5 and 0.7 correspond to the maximum
sensitivities Ms = 1.99, 1.47 and 1.28 respectively.
1000 1000a2
P1 (s) = , P2 (s) =
s+1 (s + 1)(s + a)2
have very similar open loop responses for large values of a. This is il-
lustrated in Figure 5.17 which shows the step responses of for a = 100.
223
Chapter 5. Feedback Fundamentals
1000
800
600
400
200
0
0 1 2 3 4 5 6 7 8
Figure 5.17 Step responses for systems with the transfer functions P1 ( s) =
1000/( s + 1) and P2 ( s) = 107 /(( s + 1)( s + 100)2 ).
The differences between the step responses are barely noticeable in the
figure. The transfer functions from reference values to output for closed
loop systems obtained with error feedback with C = 1 are
1000 107
T1 = , T2 =
s + 1001 (s − 287)(s2 + 86s + 34879)
The closed loop systems are very different because the system T1 is stable
and T2 is unstable.
1000 1000
P1 (s) = , P2 (s) =
s+1 s−1
have very different open loop properties because one system is unstable
and the other is stable. The transfer functions from reference values to
output for closed loop systems obtained with error feedback with C = 1
are
1000 1000
T1(s) = T2(s) =
s + 1001 s + 999
which are very close.
224
5.7 The Sensitivity Functions
P1 C P2 C ( P1 − P2 ) C
δ ( P1 , P2 ) = − = (5.18)
1 + P1 C 1 + P2 C (1 + P1 C)(1 + P2 C)
P1 − P2
δ ( P1 , P2 )
P1 P2 C
δ ( P1 , P2 ) Ms1 Ms2h C( P1 − P2 )h
For frequencies where P1 and P2 have small gains, typically for high
frequencies, we have
δ ( P1 , P2 ) h C( P1 − P2 )h
We have seen that the sensitivity function S and the complementary sen-
sitivity function T tell much about the feedback loop. We have also seen
from Equations (5.6) and (5.14) that it is advantageous to have a small
value of the sensitivity function and it follows from (5.11) that a small
value of the complementary sensitivity allows large process uncertainty.
Since
1 P(s) C(s)
S(s) = and T (s) =
1 + P(s) C(s) 1 + P(s) C(s)
225
Chapter 5. Feedback Fundamentals
it follows that
S(s) + T (s) = 1 (5.19)
This means that S and T cannot be made small simultaneously. The loop
transfer function L is typically large for small values of s and it goes to
zero as s goes to infinity. This means that S is typically small for small s
and close to 1 for large. The complementary sensitivity function is close
to 1 for small s and it goes to 0 as s goes to infinity.
A basic problem is to investigate if S can be made small over a large
frequency range. We will start by investigating an example.
k
L(s) = P(s) C(s) =
s+1
where parameter k is the controller gain. The sensitivity function is
s+1
S(s) =
s+1+ k
and we have r
1 + ω2
h S(iω )h =
1 + 2k + k2 + ω 2
This implies that h S(iω )h < 1 for all finite frequencies and that the sensi-
tivity can be made arbitrary small for any finite frequency by making k
sufficiently large.
The system in Example 5.6 is unfortunately an exception. The key feature
of the system is that the Nyquist curve of the process lies in the fourth
quadrant. Systems whose Nyquist curves are in the first and fourth quad-
rant are called positive real. For such systems the Nyquist curve never
enters the region shown in Figure 5.11 where the sensitivity is greater
than one.
For typical control systems there are unfortunately severe constraints
on the sensitivity function. Bode has shown that if the loop transfer has
poles pk in the right half plane and if it goes to zero faster than 1/s for
large s the sensitivity function satisfies the following integral
Z ∞ Z ∞ X
1
log h S(iω )hdω = log dω = π Re pk (5.20)
0 0 h1 + L(iω )h
This equation shows that if the sensitivity function is made smaller for
some frequencies it must increase at other frequencies. This means that
226
5.7 The Sensitivity Functions
0
log hS(iω )h
−1
−2
PSfrag replacements
−3
0 0.5 1 1.5 2 2.5 3
ω
= I1 + I2 + I3 = 0
227
Chapter 5. Feedback Fundamentals
pk
PSfrag replacements
where R is a large semi circle on the right and γ k is the contour starting
on the imaginary axis at s = Im p k and a small circle enclosing the pole
pk. The integral is zero because the function log S(s) is regular inside the
contour. We have
Z iR Z iR
I1 = − i log( S(iω ))dω = −2i log(h S(iω )h)dω
− iR 0
because the real part of log S(iω ) is an even function and the imaginary
part is an odd function. Furthermore we have
Z Z Z
I2 = log( S(s))ds = log(1 + L(s))ds L(s)ds
R R R
Since L(s) goes to zero faster than 1/s for large s the integral goes to
zero when the radius of the circle goes to infinity. Next we consider the
integral I3, for this purpose we split the contour into three parts X + , γ
and X − as indicated in Figure 5.19. We have
Z Z Z Z
log S(s)ds = log S(s)ds + log S(s)ds + log S(s)ds
γ X+ γ X−
The contour γ is a small circle with radius r around the pole p k. The
magnitude of the integrand is of the order log r and the length of the
path is 2π r. The integral thus goes to zero as the radius r goes to zero.
228
5.8 Reference Signals
Furthermore we have
Z Z
log S(s)ds + log S(s)ds
X+ X−
Z
= log S(s) − log S(s − 2π i ds = 2π pk
X+
Letting the small circles go to zero and the large circle go to infinity and
adding the contributions from all right half plane poles p k gives
Z iR X
I1 + I2 + I3 = −2i log h S(iω )hdω + 2π pk = 0.
0 k
PCF
G yr = = FT
1 + PC
CF
Gur =
1 + PC
First we can observe that if F = 1 then the response to reference signals
is given by T. In many cases the transfer function T gives a satisfactory
response but in some cases it may be desirable to modify the response. If
the feedback controller C has been chosen to deal with disturbances and
process uncertainty it is straight forward to find a feedforward transfer
function that gives the desired response. If the desired response from
reference r to output y is characterized by the transfer function M the
transfer function F is simply given by
M (1 + PC) M
F= = (5.21)
T PC
The transfer function F has to be stable and it therefore follows that all
right half plane zeros of C and P must be zeros of M . Non-minimum phase
properties of the process and the controller therefore impose restrictions
on the response to reference signals. The transfer function given by (5.21)
can also be complicated so it may be useful to approximate the transfer
function.
229
Chapter 5. Feedback Fundamentals
S(s) = e0 + e1s + e2 s2 + . . . .
dr(t) d2 r(t)
y(t) = r(t) − e0 r(t) − e1 − e2 +... (5.22)
dt dt2
1
S(s) =
1 + P(s) C(s)
The coefficient e0 is thus zero if P(s) C(s) 1/s for small s, i.e. if the
process or the controller has integral action.
r(t) = v0 t.
230
5.8 Reference Signals
F = 1 + f 1s (5.23)
1
E(s) = S(s)
s
E(s) 1
= S(s) 2
s s
231
Chapter 5. Feedback Fundamentals
1.4
1.2
Step response
1
0.8
0.6
0.4
0.2
0
0 1 2 3 4 5 6 7 8 9 10
Time
10
Ramp response
4
PSfrag replacements
2
0
0 1 2 3 4 5 6 7 8 9 10
Time
Figure 5.20 Step (above) and ramp (below) responses for a system with error
feedback having e0 = e1 = 0.
Since the integral of the error for a step in the reference is zero it means
that the error must have an overshoot. This is illustrated in Figure 5.20.
This is avoided if feedforward is used.
The figure indicates that an attempt to obtain a controller that gives
good responses to step and ramp inputs is a difficult compromise if the
controller is linear and time invariant. In this case it is possible to resolve
the compromise by using adaptive controllers that adapt their behavior
to the properties of the input signal.
Constraints on the sensitivities will, in general, give restrictions on the
closed loop poles that can be chosen. This implies that when controllers
are designed using pole placements it is necessary to check afterwards
that the sensitivities have reasonable values. This does in fact apply to all
design methods that do not introduce explicit constraints on the sensitivity
functions.
Analysis of the sensitivity functions has shown that the closed loop poles
must be related to the poles and zeros of the process in order to have a
232
5.9 Implications for Pole Placement Design*
A Design Procedure
To preform a design we start with a mathematical model of the system in
the form of the transfer function
n(s) −sT
P(s) = e
d(s)
Since the design procedure is based on the assumption that the transfer
function is rational the time delay is approximated by
1 − sT /2
e−sT
1 + sT /2
This gives a process model having the transfer function
n(s)(1 − sT /2) b(s)
P(s) = = (5.24)
d(s)(1 + sT /2) a(s)
With this approximation a time delay T shows up as a right half plane
zero at s = 2/T. The controller is assumed to have the transfer function
g(s)
C(s) =
f (s)
The closed loop characteristic polynomial is
233
Chapter 5. Feedback Fundamentals
b
P(s) =
s+a
the closed loop system is completely characterized by two poles. In Exam-
ple 5.3 it was shown that if ω 0 < a it is necessary to choose the closed
loop poles as
c(s) = (s + ω 0 )(s + a)
This means that there is a single dominant pole ω 0 and that the fast
process pole appears as a pole of the closed loop system. The controller is
ω 0 (s + a)
C(s) =
bs
which implies that the fast process pole s = −a is canceled by a controller
zero. The closed loop system will approximately behave like a first order
system with the pole s = −ω 0 . The response time of the closed loop system
is approximately 1/ω 0, which is slower than the response time of the open
loop system. If a faster response is required, i.e. ω 0 ≥ a, the polynomial
can be chosen as
c(s) = s2 + 2ζ 0ω 0 s + ω 20
The controller is then
(2ζ 0ω 0 − a)s + ω 20
C(s) =
bs
P
P
1 + PC
This means that the closed loop system has poles close to process poles
larger than ω 0 . This implies that ω 0 must be larger than the h p rhph, where
prhp is the process pole in the right half plane with the largest magnitude.
When the loop gain is large we have
C 1
1 + PC P
234
5.9 Implications for Pole Placement Design*
This means that the closed loop system has poles close to the process zeros
smaller than ω 0 . This implies that ω 0 must be smaller than the h z rhph,
where zrhp is the process zer in the right half plane with the smallest
magnitude.
Process poles and zeros in the right half plane thus give severe restric-
tions in the choice of dominant poles.
1 a(s) f (s)
S= =
1 + PC a(s) f (s) + b(s)g(s)
PC b(s)g(s)
T= =
1 + PC a(s) f (s) + b(s)g(s)
235
Chapter 5. Feedback Fundamentals
– Right half plane process poles imply that ω 0 > h prhph, where prhp
is the right half plane process pole with largest magnitude.
– Right half plane process zeros imply that ω 0 < h zrhph, where prhp
is the right half plane process zero with smallest magnitude.
– Notice that there are process with poles and zeros in the right
half plane such that it may not be possible to find ω 0 . The
process must then be modified by adding actuators and sensors
or by redesign.
Real poles and zeros can be neglected if their total phase lag at
ω 0 is less than 0.05-0.2 rad.
Complex poles can be neglected if the loop transfer function at
the resonance has magnitude less than 1/ M s. This guarantees
that they will not give a large sensitivity even if there is a large
phase error.
• Cancel fast stable process poles and slow stable process zeros by
correspondign controller poles and zeros.
• Solve the pole placement problem for the reduced process model.
• Obtain the complete controller by adding the canceled process poles
and zeros.
236
5.9 Implications for Pole Placement Design*
237
Chapter 5. Feedback Fundamentals
c(s) = c̃(s)ā(s)b(s)
fˆ(s) = f˜ (s)b(s)
g(s) = g̃(s)ā(s)
where
sa(s) f˜ (s) + b̄(s)g̃(s) = c̃(s) (5.27)
An Example
The design procedure will be illustrated by an few examples. Consider a
process with the transfer function
1
P(s) = (5.28)
(s + 1)(0.1s + 1)(0.01s + 1)(0.001s + 1)
P(s) 1 (5.29)
The total phase lag of the neglected process poles at ω 0 is 1.111ω 0. Using
the design rules this means that ω 0 has to be slightly smaller than 0.2.
The approximated process can be controlled by the integrating controller
ω0 ki
C(s) = =
s s
238
5.9 Implications for Pole Placement Design*
Figure 5.21 Step response of the gang of four for integrating control of the process
P( s) = 1/(( s + 1)(0.1s + 1)(0.01s + 1)(0.001s + 1)) with k i = 0.1 (full), 0.2 (dashed)
and 0.5(dash-dotted). The dotted curves shows the response for the PI controllers
with the simplified process model P( s) = 1.
which gives a closed loop system of first order with the pole s = ki = ω 0 .
The closed loop characteristic polynomial with the true model is
The closed loop system has one pole at -0.114 which is close to the specified
value 0.1. The next pole at -0.87 is close to the open loop pole at -1 and the
remaining poles are very close to the open loop poles at -10, -100 and -1000.
Figure 5.21 shows the step responses of the Gang of Four. The responses
PC/(1 + PC), C/(1 + PC) and PC/(1 + PC) are close to responses of first
order systems for ω 0 = 0.1 and 0.2. There are substantial deviations for
ω 0 = ki = 0.5 and even closer for ω 0 = 0.1. The higher order dynamics is
clearly visible in the step of responses P/(1 + PC).
239
Chapter 5. Feedback Fundamentals
y11 y12
0.6
1
0.5
0.8
0.4
0.6
0.3
0.4 0.2
0.2 0.1
0 0
0 2 4 6 8 10 0 2 4 6 8 10
y21 y22
2.5 1
2 0.8
0.6
1.5
PSfrag replacements 0.4
1
0.2
0.5
0
0
0 2 4 6 8 10 0 2 4 6 8 10
Figure 5.22 Step response of the gang of four for PI control of the process P( s) =
1/(( s + 1)(0.1s + 1)(0.01s + 1)(0.001s + 1)) with the controller C ( s) = k i (1 + 1/ s)
for ki = 0.5 (full), 1 (dashed) and 2 (dash-dotted). The dotted curves shows the
response for the PI controller with the simplified process model P( s) = 1/( s + 1).
1
P(s) . (5.30)
s+1
for 0.1 < ω s < 1 we simply cancel the process pole at s = 1 with a
controller zero. After cancellation the process transfer function becomes
P(s) = 1 and we can use integrating controller. With the approximated
model the closed loop system is still of first order with the dominating
pole ω 0 = ki . Adding the controller zero required to cancel the process
pole the complete controller becomes
s+1
C(s) = ki
s
With k = 1 the closed loop system has poles at s = −1.0, -1.14, -8.7, -
100 and -1000. Figure 5.22 shows the step responses of the Gang of Four.
The responses y11 and y21 are close to responses of first order systems for
all values of ki in the figure. To get a significant deveiation ki has to be
240
5.9 Implications for Pole Placement Design*
(2ζ 0ω 0 − 1)s + ω 20
C(s) =
s
PID Control A comparison of Figures 5.21 and 5.22 shows that the
response speed can be increased significantly by using PI control instead
of an pure integrating control. The response speed can be increased even
higher by using PID control. Since a PID controller permits arbitrary pole
placement for a second order system the process will be approximated by
1 10
P(s) = (5.31)
(s + 1)(0.1s + 1) (s + 1)(s + 10)
The phase lag at ω 0 of the neglected poles is 0.011ω 0 which implies that
the model permits values of ω 0 up to 20. For ω 0 < 10 the controller can
be designed simply by canceling the fast process pole at s = −10. This
gives the process model P1 (s) = 1/(s + 1). Equation (5.27) then becomes
There is a solution with deg g̃(s) = 1 deg f̂(s) = 1 which means that
deg c̃(s) = 3. We choose two complex dominant poles specified by ω 0 and
ζ 0 . The remaining closed loop pole is chosen as s = − Nω 0 . Hence
1 + f 1 = (2ζ 0 + N )ω 0
f 1 + g0 = (1 + 2ζ 0 N )ω 20
g1 = Nω 30
f 1 = (2ζ 0 + N )ω 0 − 1
g0 = (1 + 2ζ 0 N )ω 20 − (2ζ 0 + N )ω 0 + 1
g1 = Nω 30
241
Chapter 5. Feedback Fundamentals
where
With ω 0 = 10 and ζ 0 = 0.7 the closed loop system has poles at s = −10 ,
−7.5 ± 8i, -58.7 -140 and -1000. These values can be compared with the
poles −7 ± 7.14 and -100 for the closed loop obtained with the simplified
model. The dominant poles are close but the real poles is different. Fig-
ure 5.23 shows the step responses of the Gang of Four. The curves indicate
that the approximations are reasonable.
242
5.10 Fundamental Limitations*
y11 y12
1.2 0.1
1 0.08
0.8
0.06
0.6
0.04
0.4
0.2 0.02
0 0
0 0.5 1 1.5 0 0.5 1 1.5
y21 y22
10 1
0.8
5
0.6
0.2
−5
0
−10 −0.2
0 0.5 1 1.5 0 0.5 1 1.5
Figure 5.23 Step response of the gang of four for PI control of the process P( s) =
1/(( s + 1)(0.1s + 1)(0.01s + 1)(0.001s + 1)) with a PID controller (5.32) designed for
ζ 0 = 0.7, ω 0 = 5 (full) and 10 (dashed). The dotted curves shows the response for
PI controllers with the simplified process model P( s) = 10/(( s + 1)( s + 10)).
control. The maximum error drops from 0.84 to 0.39. The improvement
is obtained by a barely perceptible increase of the high frequency gain of
the controller. The response time also decreases significantly as is seen
from the values of the integral gain and the step responses.
Response time is decreased even further by using PID control. Load
disturbance attenuation also improves substantially, e max decreases from
0.39 to 0.038 compared to PID control. A large increase of controller gain
is however required to obtain this improvement.
• Measurement noise
243
Chapter 5. Feedback Fundamentals
• Actuator saturation
• Process dynamics
Dynamics Limitations
The limitations caused by noise and saturations seem quite obvious. It
turns out that there may also be severe limitations due to the dynamical
properties of the system. This means that there are systems that are in-
herently difficult or even impossible to control. Designers of any system
should be aware of this. Since systems are often designed from static con-
siderations the difficulties caused by dynamics do not show up. We have
already encountered this in Section 5.9 where we found that it was nec-
essary to choose the dominant pole so that ω 0 is larger than the fastest
unstable pole but smaller than the slowest zero. It seems intuitively rea-
sonable that a fast closed loop system is required to stabilize an unstable
pole and that the response speed should be matched to the unstable pole.
In the same way it seems reasonable that it is not possible to get a very
rapid response for a system with a time delay. Since a right half plane
zero s = z is similar to a time delay T = 1/2z it then follows that a right
half plane zero limits the achievable response time.
244
5.10 Fundamental Limitations*
• A right half plane zero z rhp limits the response speed. A simple rule
of thumb is
ω g c < 0.5zrhp (5.34)
Slow RHP zeros are thus particularly bad. Notice that if a time delay
is approximated by a zero in the right half plane we can apply the
rule for right half plane zeros to get ω g c Td < 1.
• A right half plane pole p rhp requires high gain crossover frequency.
A simple rule of thumb is
• Systems with a right half plane pole p and a right half plane zero z
cannot be controlled unless the pole and the zero are well separated.
A simple rule of thumb is
• A system with a a right half plane pole and a time delay Td cannot be
controlled unless the product p rhpTd is sufficiently small. A simple
rule of thumb is
prhpTd < 0.16 (5.37)
245
Chapter 5. Feedback Fundamentals
s2
G (s) =
s2 − g
Q
mV0 Q Js2 − mg l
P(s) =
b −as + V0
Notice that the transfer function depends strongly on the forwardpvelocity
of the bicycle. The system thus has a right half plane pole at p = mgQ/ J
and a right half plane zero at z = V0/a, and it can be suspected that the
system is difficult to control. The location of the pole does not depend on
velocity but the the position of the zero changes significantly
p with velocity.
At low velocities the zero is at the origin. For V0 = a mgQ/ J the pole and
the zero are at the same location and for higher velocities the zero is to
the right of the pole. To draw some quantitative conclusions we introduce
the numerical values m = 70 kg, Q = 1.2 m, a = 0.7, J = 120 kgm2 and
V = 5 m/s, give z = V /a = 7.14 rad/s and p = ω 0 = 2.6 rad/s we find
that p = 2.6. With V0 = 5 m/s we get z = 7.1, and p/ z = 2.7. To have a
situation where the system can be controlled it follows from (5.36) that
to have z/ p = 6 the velocity must be increased to 11 m/s. We can thus
conclude that if the speed of the bicycle can be increased to about 10 m/s
so rapidly that we do not loose balance it can indeed be ridden.
The bicycle example illustrates clearly that it is useful to assess the funda-
mental dynamical limitations of a system at an early stage in the design.
If this had been done the it could quickly have been concluded that the
study of rear wheel steered motor bikes in 4.6 was not necessary.
246
5.11 Electronic Amplifiers
Remedies
Having understood factors that cause fundamental limitations it is inter-
esting to know how they should be overcome. Here are a few suggestions.
Problems with sensor noise are best approached by finding the roots
of the noise and trying to eliminate them. Increasing the resolution of
a converter is one example. Actuation problems can be dealt with in a
similar manner. Limitations caused by rate saturation can be reduced by
replacing the actuator.
Problems that are caused by time delays and RHP zeros can be ap-
proached by moving sensors to different places. It can also be beneficial to
add sensors. Recall that the zeros depend on how inputs and outputs are
coupled to the states of a system. A system where all states are measured
has no zeros.
Poles are inherent properties of a system, they can only be modified
by redesign of the system.
Redesign of the process is the final remedy. Since static analysis can
never reveal the fundamental limitations it is very important to make an
assessment of the dynamics of a system at an early stage of the design.
This is one of the main reasons why all system designers should have a
basic knowledge of control.
247
PSfrag replacements
V1 V V2
R2
R1 + R 2 Σ −A
R1
R1 + R 2
PSfrag replacements
Figure 5.24 Block diagram of the feedback amplifier in Figure 2.9. Compare with
Figure 2.10
V1 R2 V V2
R1 + R 2 Σ −A Σ
R1
R1 + R 2
of the feedback loop in the forward path and the standard −1 block has
been replaced by a feedback. It is customary to use diagrams of this type
when dealing with feedback amplifiers. The generic version of the diagram
is shown in Figure 5.25. The block A represents the open loop amplifier,
block F the feedback and block H the feedforward. The blocks F and H
are represented by passive components.
For the circuit in Figure 5.24 we have
R1
F=
R1 + R 2
R2
H=
R1 + R 2
L V2
= −G
L V1
where
AH
G= (5.38)
1 + AF
248
5.11 Electronic Amplifiers
b
A(s) =
s+a
The amplifier has gain b/a and bandwidth a. The gain bandwidth product
is b. Typical numbers for a simple operational amplifier that is often used
to implement control systems, LM 741, are a = 50 Hz, b = 1 MHz. Other
amplifiers may have gain bandwidth products that are several orders of
magnitude larger.
Furthermore we have
R1 R2
F= , H=
R1 + R 2 R1 + R 2
bR2 bR2
G=
( R1 + R2 )(s + a) + bR1 R2 s + bR1
Gain B andwidth = b
Notice that feedback does not change the gain bandwidth product. The
effect of feedback is simply to decrease the gain and increase the band-
width. This is illustrated in Figure 5.26 which shows the gain curves of
the open and closed loop systems. Also notice that the sensitivity of the
system is
1 ( R1 + R2 )(s + a) R2 (s + a)
S= =
1 + AF ( R1 + R2 )(s + a) + bR1 R2 s + bR1
The high open loop gain of the amplifier is traded off for high bandwidth
and low sensitivity. This is some times expressed by saying that gain is the
hard currency of feedback amplifiers which can be traded for sensitivity
and linearity.
Sensitivity
It follows from (5.38) that
249
Chapter 5. Feedback Fundamentals
6
10
4
10
Gain
2
10
frag replacements 0
10
−2 −1 0 1 2 3 4 5 6
10 10 10 10 10 10 10 10 10
ω
Figure 5.26 Gain curves of the open loop amplifier (full lines) and the feedback
amplifier (dashed)
d log G 1
=
d log A 1 + AF
d log G AF
=−
d log F 1 + AF
d log G
=1
d log H
The loop transfer function is normally large which implies that it is only
the sensitivity with respect the amplifier that is small. This is, how-
ever, the important active part where there are significant variations.
The transfer functions F and H typically represent passive components
that are much more stable than the amplifiers.
Vol = N2 − A( N1 + H V1)
250
5.12 Summary
The signals will be smaller for a system with feedback but the signal to
noise ratio does not change.
5.12 Summary
Having got insight into some fundamental properties of the feedback loop
we are in a position to discuss how to formulate specifications on a control
system. It was mentioned in Section 5.2 that requirements on a control
system should include stability of the closed loop system, robustness to
model uncertainty, attenuation of measurement noise, injection of mea-
surement noise ability to follow reference signals. From the results given
in this section we also know that these properties are captured by six
transfer functions called the Gang of Six. The specifications can thus be
expressed in terms of these transfer functions.
Stability and robustness to process uncertainties can be expressed by
the sensitivity function and the complementary sensitivity function
1 PC
S= , T= .
1 + PC 1 + PC
P
G yd = = PS.
1 + PC
C
− Gun = = CS,
1 + PC
which describes how measurement noise influences the control signal. The
response to set point changes is described by the transfer functions
FPC FC
G yr = = FT , Gur = = FCS
1 + PC 1 + PC
251
Chapter 5. Feedback Fundamentals
252
6
PID Control
6.1 Introduction
The PID controller is the most common form of feedback. It was an es-
sential element of early governors and it became the standard tool when
process control emerged in the 1940s. In process control today, more than
95% of the control loops are of PID type, most loops are actually PI con-
trol. PID controllers are today found in all areas where control is used.
The controllers come in many different forms. There are stand-alone sys-
tems in boxes for one or a few loops, which are manufactured by the
hundred thousands yearly. PID control is an important ingredient of a
distributed control system. The controllers are also embedded in many
special-purpose control systems. PID control is often combined with logic,
sequential functions, selectors, and simple function blocks to build the
complicated automation systems used for energy production, transporta-
tion, and manufacturing. Many sophisticated control strategies, such as
model predictive control, are also organized hierarchically. PID control is
used at the lowest level; the multivariable controller gives the setpoints
to the controllers at the lower level. The PID controller can thus be said
to be the “bread and buttert’t’ of control engineering. It is an important
component in every control engineer’s tool box.
PID controllers have survived many changes in technology, from me-
chanics and pneumatics to microprocessors via electronic tubes, transis-
tors, integrated circuits. The microprocessor has had a dramatic influence
on the PID controller. Practically all PID controllers made today are based
on microprocessors. This has given opportunities to provide additional fea-
tures like automatic tuning, gain scheduling, and continuous adaptation.
253
Chapter 6. PID Control
We will start by summarizing the key features of the PID controller. The
“textbook” version of the PID algorithm is described by:
Zt
1 de(t)
u(t) = K e(t) + e(τ )dτ + Td (6.1)
Ti dt
0
254
6.2 The Algorithm
K =5
K =2
1
K =1
0
0 5 10 15 20
K =5
K =2
2 K =1
−2
0 5 10 15 20
Figure 6.1 Simulation of a closed-loop system with proportional control. The pro-
cess transfer function is P( s) = 1/( s + 1)3 .
Ti = 1
Ti = 2
Ti = 5
1
Ti = ∞
0
0 5 10 15 20
Ti = 1
2 Ti = 2
Ti = 5
1 Ti = ∞
0
0 5 10 15 20
Figure 6.2 Simulation of a closed-loop system with proportional and integral con-
trol. The process transfer function is P( s) = 1/( s + 1)3 , and the controller gain is
K = 1.
A Perspective
There is much more to PID than is revealed by (6.1). A faithful imple-
mentation of the equation will actually not result in a good controller. To
obtain a good PID controller it is also necessary to consider
255
Chapter 6. PID Control
Td = 0.1
Td = 0.7
1
Td = 4.5
0
0 5 10 15 20
−2
0 5 10 15 20
• Windup
• Tuning
• Computer implementation
In the case of the PID controller these issues emerged organically as the
technology developed but they are actually important in the implemen-
tation of all controllers. Many of these questions are closely related to
fundamental properties of feedback, some of them have been discussed
earlier in the book.
Filtering
Differentiation is always sensitive to noise. This is clearly seen from the
transfer function G (s) = s of a differentiator which goes to infinity for
large s. The following example is also illuminating.
256
6.3 Filtering and Set Point Weighting
d y(t)
= cos t + n(t) = cos t + a nω cos ω n t
dt
The signal to noise ratio for the original signal is 1/a n but the signal to
noise ratio of the differentiated signal is ω /a n . This ratio can be arbitrarily
high if ω is large.
In a practical controller with derivative action it is therefor necessary to
limit the high frequency gain of the derivative term. This can be done by
implementing the derivative term as
sK Td
D=− Y (6.2)
1 + sTd / N
lim C(s) = − K (1 + N )
s→∞
257
Chapter 6. PID Control
1
F (s) =
(1 + sT f )n
where T f is the filter time constant and n is the order of the filter. The
choice of T f is a compromise between filtering capacity and performance.
The value of T f can be coupled to the controller time constants in the
same way as for the derivative filter above. If the derivative time is used,
T f = Td/ N is a suitable choice. If the controller is only PI, T f = Ti / N
may be suitable.
The controller can also be implemented as
1 1
C(s) = − K 1 + + sTd (6.3)
sTi (1 + sTd / N )2
This structure has the advantage that we can develop the design meth-
ods for an ideal PID controller and use an iterative design procedure. The
controller is first designed for the process P(s). The design gives the con-
troller parameter Td . An ideal controller for the process P(s)/(1 + sTd/ N )2
is then designed giving a new value of Td etc. Such a procedure will also
give a clear picture of the trade-off between performance and filtering.
Zt
1 dr(t) d y(t)
u(t) = K br(t) − y(t) + e(τ )dτ + Td c − (6.4)
Ti dt dt
0
where b and c are additional parameter. The integral term must be based
on error feedback to ensure the desired steady state. The controller given
by (6.4) has a structure with two degrees of freedom because the signal
path from y to u is different from that from r to u. The transfer function
from r to u is
U (s) 1
= Cr (s) = K b + + csTd (6.5)
R(s) sTi
258
6.4 Different Parameterizations
1.5
Output y
1
0.5
0
0 5 10 15
4
Input u
ag replacements 2
0
0 5 10 15
Time t
Figure 6.4 Response to a step in the reference for systems with different set point
weights b = 0 dashed, b = 0.5 full and b = 1.0 dash dotted. The process has the
transfer function P( s) = 1/( s + 1)3 and the controller parameters are k = 3, ki = 1.5
and kd = 1.5.
Set point weighting is thus a special case of controllers having two degrees
of freedom.
The system obtained with the controller (6.4) respond to load distur-
bances and measurement noise in the same way as the controller (6.1)
. The response to reference values can be modified by the parameters b
and c. This is illustrated in Figure 6.4, which shows the response of a PID
controller to setpoint changes, load disturbances, and measurement errors
for different values of b. The figure shows clearly the effect of changing b.
The overshoot for setpoint changes is smallest for b = 0, which is the case
where the reference is only introduced in the integral term, and increases
with increasing b.
The parameter c is normally zero to avoid large transients in the con-
trol signal due to sudden changes in the setpoint.
259
Chapter 6. PID Control
transfer function
1
G (s) = K 1+ + sTd (6.7)
sTi
A slightly different version is most common in commercial controllers.
This controller is described by
1 TdP 1
G P (s) = K P 1 + (1 + sTd
P
) = K P
1 + + + sTd
P
(6.8)
sTiP TiP sTiP
TiP + TdP
K = KP
TiP
Ti = TiP + TdP (6.9)
TiP TdP
Td =
TiP + TdP
Ti ≥ 4Td
K p
KP = 1+ 1 − 4Td/Ti
2
T p
i
TiP = 1+ 1 − 4Td/Ti (6.10)
2
Ti p
TdP = 1− 1 − 4Td/Ti
2
260
6.4 Different Parameterizations
when both the I and the D parts of the controller are used. If we only use
the controller as a P, PI, or PD controller, the two forms are equivalent.
Yet another representation of the PID algorithm is given by
ki
G PP (s) = k + + skd (6.11)
s
K
k= K ki = kd = K Td
Ti
kPi
G PI (s) = kP +
s
G PD (s) = 1 + kPd s
Notice that the proportional gain of the PD controller must be one in order
to have zero steady state error. The input-output relation of the complete
controller is
kPi
U (s) = kP R(s) + ( R(s) − Y (s)) − ( kP + kPd hPi) Y (s) − kP kPd sY (s)
s
Which shows that the controller is thus identical to the controller given
261
Chapter 6. PID Control
d
PSfrag replacements
r u y
Σ PI Σ Process
PSfrag replacements
PD
r e y
Σ PI Σ P
PD
−1
d
k = kP + kPd kPi
ki = kPi
kd = kP kPd
kP + kPd kPi kP
Ti = =
kPi kPi
kP kPd
Td =
kPi
kP
b=
kP + kPd hPi
c=0
262
6.5 Windup
uncertainty. When this is done the response to set points can be adjusted
by choosing the parameters b and c. The controller parameters appear in
a much more complicated way in the PIPD controller.
6.5 Windup
263
Chapter 6. PID Control
2 y
1
ysp
0
0 20 40 60 80
0.1
u
−0.1
0 20 40 60 80
I
2
−2
0 20 40 60 80
Figure 6.6 Illustration of integrator windup. The diagrams show process output
y, setpoint ysp , control signal u, and integral part I.
under labels like preloading, batch unit, etc. Although the problem was
well understood, there were often restrictions caused by the analog tech-
nology. The ideas were often kept as trade secrets and not much spoken
about. The problem of windup was rediscovered when controllers were im-
plemented digitally and several methods to avoid windup were presented
in the literature. In the following section we describe some of the methods
used to avoid windup.
Setpoint Limitation
One attempt to avoid integrator windup is to introduce limiters on the
setpoint variations so that the controller output never reaches the actua-
tor limits. This frequently leads to conservative bounds and poor perfor-
mance. Furthermore, it does not avoid windup caused by disturbances.
Incremental Algorithms
In the early phases of feedback control, integral action was integrated
with the actuator by having a motor drive the valve directly. In this
case windup is handled automatically because integration stops when the
valve stops. When controllers were implemented by analog techniques,
264
6.5 Windup
–y
KT d s
Actuator
model Actuator
e = r− y
K Σ
K 1 – +
Ti
Σ Σ
s
1 es
Tt
Figure 6.7 Controller with anti-windup where the actuator output is estimated
from a mathematical model.
265
Chapter 6. PID Control
ysp
1
y
0.5
0
0 10 20 30
0.15
u
0.05
−0.05
0 10 20 30
0
I
−0.4
−0.8
0 10 20 30
Figure 6.8 Controller with anti-windup applied to the system of Figure 6.6. The
diagrams show process output y, setpoint ysp , control signal u, and integral part I.
process input remains constant. There is, however, a feedback path around
the integrator. Because of this, the integrator output is driven towards a
value such that the integrator input becomes zero. The integrator input
is
1 K
es + e
Tt Ti
where e is the control error. Hence,
K Tt
es = − e
Ti
K Tt
v = ulim + e
Ti
where ulim is the saturating value of the control variable. This means that
the signal v settles on a value slightly out side the saturation limit and the
control signal can react as soon as the error changes time. This prevents
266
6.5 Windup
ysp Tt = 3
Tt = 2
1
Tt = 0.1, Tt = 1
0
0 10 20 30
0.1 Tt = 3
Tt = 0.1 Tt = 2
0 Tt = 1
−0.1
0 10 20 30
Figure 6.9 The step response of the system in Figure 6.6 for different values of
the tracking time constant Tt . The upper curve shows process output y and setpoint
ysp , and the lower curve shows control signal u.
the integrator from winding up. The rate at which the controller output is
reset is governed by the feedback gain, 1/Tt, where Tt can be interpreted
as the time constant, which determines how quickly the integral is reset.
We call this the tracking time constant.
It frequently happens that the actuator output cannot be measured.
The anti-windup scheme just described can be used by incorporating a
mathematical model of the saturating actuator, as is illustrated in Fig-
ure 6.7.
Figure 6.8 shows what happens when a controller with anti-windup is
applied to the system simulated in Figure 6.6. Notice that the output of
the integrator is quickly reset to a value such that the controller output
is at the saturation limit, and the integral has a negative value during
the initial phase when the actuator is saturated. This behavior is drasti-
cally different from that in Figure 6.6, where the integral has a positive
value during the initial transient. Also notice the drastic improvement in
performance compared to the ordinary PI controller used in Figure 6.6.
The effect of changing the values of the tracking time constant is il-
lustrated in Figure 6.9. From this figure, it may thus seem advantageous
to always choose a very small value of the time constant because the
integrator is then reset quickly. However, some care must be exercised
when introducing anti-windup in systems with derivative action. If the
time constant is chosen too small, spurious errors can cause saturation
of the output, which accidentally resets the integrator. The tracking time
constant Tt should be larger than Td and smaller than Ti . A rule of thumb
267
Chapter 6. PID Control
y sp P
b Σ K
y sKTd D
−1 Σ
1+ sTd / N
e K 1 I v
Ti Σ s
Σ
1
Tt
w + –
Σ
y sp
SP v
y MV PID
w TR
Figure 6.10 Block diagram and simplified representation of PID module with
tracking signal.
√
that has been suggested is to choose Tt = Ti Td.
268
6.6 Tuning
A
SP
MV PID Actuator
TR
B
Actuator model
SP
v u
MV PID Actuator
TR
Figure 6.11 Representation of the controllers with anti-windup in Figure 6.7 us-
ing the basic control module with tracking shown in Figure 6.10.
6.6 Tuning
All general methods for control design can be applied to PID control. A
number of special methods that are tailor-made for PID control have also
been developed, these methods are often called tuning methods. Irrespec-
tive of the method used it is essential to always consider the key elements
of control, load disturbances, sensor noise, process uncertainty and refer-
ence signals.
The most well known tuning methods are those developed by Ziegler
and Nichols. They have had a major influence on the practice of PID
control for more than half a century. The methods are based on character-
ization of process dynamics by a few parameters and simple equations for
the controller parameters. It is surprising that the methods are so widely
referenced because they give moderately good tuning only in restricted
situations. Plausible explanations may be the simplicity of the methods
and the fact that they can be used for simple student exercises in basic
control courses.
269
Chapter 6. PID Control
Table 6.1 PID controller parameters obtained for the Ziegler-Nichols step response
method.
Controller K Ti Td Tp
P 1/a 4L
PI 0.9/a 3L 5.7L
PID 1.2/a 2L L/2 3.4L
Figure 6.12.
The point where the slope of the step response has its maximum is
first determined, and the tangent at this point is drawn. The intersections
between the tangent and the coordinate axes give the parameters a and L.
The controller parameters are then obtained from Table 6.1. An estimate
of the period Tp of the closed-loop system is also given in the table.
270
6.6 Tuning
0.5
−0.5
−1
−0.5 0 0.5 1
Controller K Ti Td Tp
P 0.5K u Tu
PI 0.4K u 0.8Tu 1.4Tu
PID 0.6K u 0.5Tu 0.125Tu 0.85Tu
process starts to oscillate. The gain when this occurs is K u and the period
of the oscillation is Tu . The parameters of the controller are then given by
Table 6.2. An estimate of the period Tp of the dominant dynamics of the
closed-loop system is also given in the table.
The frequency response method can be viewed as an empirical tuning
procedure where the controller parameters are obtained by direct experi-
ments on the process combined with some simple rules. For a proportional
controller the rule is simply to increase the gain until the process oscil-
lates and then reduce it by 50%.
271
Chapter 6. PID Control
kp
P1 (s) = e−sL
1 + sT
kv −sL
P2 (s) = e
s
The transfer function P1 (s), which is called a first order system with time
delay or a K LT model. Parameter L is determined from the intercept
of the tangent with largest slope with the time axis as was described in
Figure 6.12. Parameter T is also determined as shown in the figure as
the difference between the time when the step response reaches 63% of
its steady state value. Parameter kp is the static gain of the system. The
parameter kv is the largest slope of the unit step response. Parameter
L is called the apparent time delay and parameter T the apparent time
constant or the apparent lag. The adverb apparent is added to indicate
that parameters are based on approximations. The parameter
L
τ=
L+T
272
6.6 Tuning
1
P1 (s) =
(s + 1)(0.2s + 1)
1
P2 (s) =
(s + 1)4
1
P3 (s) = e−1.2s
0.05s + 1)2
The process P1 (s) has lag dominated dynamics, process P3 (s) has delay
dominated dynamics and process P2 (s) has balanced dynamics.
Figure 6.14 shows the response to a step change in the reference at
time zero and a step change in a load disturbance at the process input
for PI control of the process P1 (s). The dashed lines show the responses
obtained by the Ziegler-Nichols step response method and the full line
shows the response obtained with the improved rule which restricted the
maximum sensitivity to 1.4. The oscillatory responses to obtained by the
Ziegler-Nichols method are clearly visible in the figure which reflects the
design choice of quarter amplitude damping. The response to load distur-
bances obtained by the Ziegler-Nichols method comes at a price of poor
sensitivity. There is also a very large overshoot in the response to refer-
ence values. Figure 6.15 shows the corresponding responses for the system
P4 (s). The oscillatory character obtained with Ziegler Nichols tuning is
clearly visible. Figure 6.16 shows the response for a process that is delay
dominated. The figure shows that Ziegler-Nichols tuning performs very
poorly for this process. Overall we find that the improved tuning rules
work for a wide range of processes and that they give robust systems
with good responses.
273
Chapter 6. PID Control
a
2
1.5
1
y
0.5
0
0 1 2 3 4 5 6 7 8 9 10
t
b
15
10
5
u
−5
−10
0 1 2 3 4 5 6 7 8 9 10
t
Figure 6.14 Behavior of closed loop systems with PI controllers designed by the
Ziegler-Nichols rule (dashed) and the improved tuning rules (solid). The process
has lag dominated dynamics with the transfer function P( s) = (s+1)(01.2s+1) .
Sampling
When the controller is implemented in a computer, the analog inputs are
read and the outputs are set with a certain sampling period. This is a
drawback compared to the analog implementations, since the sampling
introduces dead-time in the control loop.
When a digital computer is used to implement a control law, the ideal
sequence of operation is the following.
1. Wait for clock interrupt
2. Read analog input
3. Compute control signal
4. Set analog output
274
6.7 Computer Implementation
a
2
1.5
1
y
0.5
0
0 10 20 30 40 50 60 70 80
t
b
2.5
1.5
1
u
0.5
−0.5
0 10 20 30 40 50 60 70 80
t
Figure 6.15 Behavior of closed loop systems with PI controllers designed by the
Ziegler-Nichols rule (dashed) and the improved tuning rules (solid). The process
has balanced dynamics with the transfer function P( s) = (s+11)4 .
Aliasing
The sampling mechanism introduces some unexpected phenomena, which
must be taken into account in a good digital implementation of a PID
controller. To explain these, consider the signals
s(t) = cos(nω st ± ω t)
and
sa (t) = cos(ω t)
275
Chapter 6. PID Control
1.5
y
0.5
0
0 2 4 6 8 10 12 14 16 18 20
t
b
1.2
0.8
0.6
u
0.4
0.2
−0.2
0 2 4 6 8 10 12 14 16 18 20
t
Figure 6.16 Behavior of closed loop systems with PI controllers designed by the
Ziegler-Nichols rule (dashed) and the improved tuning rules (solid). The process
has delay dominated dynamics with the transfer function P( s) = 0.05s1+1)2 e−1.2s .
The signals s and s a thus have the same values at the sampling instants.
This means that there is no way to separate the signals if only their
values at the sampling instants are known. Signal s a is, therefore, called
an alias of signal s. This is illustrated in Figure 6.17. A consequence of the
aliasing effect is that a high-frequency disturbance after sampling may
appear as a low-frequency signal. In Figure 6.17 the sampling period is
1 s and the sinusoidal disturbance has a period of 6/5 s. After sampling,
the disturbance appear as a sinusoid with the frequency
5
fa = 1 − = 1/6 Hz
6
276
6.7 Computer Implementation
sa
1
0 s
−1
0 1 2 3 4 5
Figure 6.17 Illustration of the aliasing effect. The diagram shows signal s and its
alias sa .
Prefiltering
The aliasing effect can create significant difficulties if proper precautions
are not taken. High frequencies, which in analog controllers normally
are effectively eliminated by low-pass filtering, may, because of aliasing,
appear as low-frequency signals in the bandwidth of the sampled control
system. To avoid these difficulties, an analog prefilter (which effectively
eliminates all signal components with frequencies above half the sampling
frequency) should be introduced. Such a filter is called an anti-aliasing
filter. A second-order Butterworth filter is a common anti-aliasing filter.
Higher-order filters are also used in critical applications. The selection of
the filter bandwidth is illustrated by the following example.
(ω s/2ω b)2 = 16
Hence,
1
ωb = ωs
8
277
Chapter 6. PID Control
Discretization
To implement a continuous-time control law, such as a PID controller in
a digital computer, it is necessary to approximate the derivatives and the
integral that appear in the control law. A few different ways to do this
are presented below.
P = K (bysp − y)
where {tk} denotes the sampling instants, i.e., the times when the com-
puter reads the analog input.
It follows that
dI K
= e (6.14)
dt Ti
The derivative is approximated by a forward difference gives
I (tk+1) − I (tk) K
= e(tk)
h Ti
This leads to the following recursive equation for the integral term
Kh
I (tk+1) = I (tk) + e(tk) (6.15)
Ti
Derivative Action The derivative term is given by Equation (6.2), i.e.
Td dD dy
+ D = − K Td (6.16)
N dt dt
This equation can be approximated in the same way as the integral
term. In this case we approximate the derivatives by a backward differ-
ence.
Td D (tk) − D (tk−1) y(tk) − y(tk−1)
+ D (tk ) = − K Td
N h h
278
6.7 Computer Implementation
Td K Td N
D (tk) = D (tk−1) − ( y(tk) − y(tk−1)) (6.17)
Td + Nh Td + Nh
Velocity Algorithms
The algorithms described so far are called positional algorithms because
the output of the algorithms is the control variable. In certain cases the
control system is arranged in such a way that the control signal is driven
directly by an integrator, e.g., a motor. It is then natural to arrange the
algorithm in such a way that it gives the velocity of the control variable.
The control variable is then obtained by integrating its velocity. An al-
gorithm of this type is called a velocity algorithm. A block diagram of a
velocity algorithm for a PID controller is shown in Figure 6.18.
Velocity algorithms were commonly used in many early controllers
that were built around motors. In several cases, the structure was re-
tained by the manufacturers when technology was changed in order to
maintain functional compatibility with older equipment. Another reason
is that many practical issues, like wind-up protection and bumpless pa-
rameter changes, are easy to implement using the velocity algorithm. This
is discussed further in Sections 6.5 and 6.7. In digital implementations
velocity algorithms are also called incremental algorithms.
Incremental algorithm
The incremental form of the PID algorithm is obtained by computing the
time differences of the controller output and adding the increments.
279
Chapter 6. PID Control
s2 KTd
ed
1+ sTd / N
External
du integrator
dt 1
ep sK Σ s
u
e K
Ti
One advantage with the incremental algorithm is that most of the com-
putations are done using increments only. Short word-length calculations
can often be used. It is only in the final stage where the increments are
added that precision is needed.
280
6.7 Computer Implementation
A
e 1 u
K s
s
B
e 1 u
K s Σ
s
+ −
Σ
+
ub
as
∆ u(t) = u(t) − u(t − h) = K e(t) + u b − u(t − h)
where h is the sampling period.
Feedforward control
In feedforward control, the control signal is composed of two terms,
u = uF B + uF F
281
Chapter 6. PID Control
the windup problem. These strategies do, however, lead to a less effective
feedforward.
Incremental algorithms are efficient for feedforward implementation.
By first adding the increments of the feedback and feedforward compo-
nents,
∆u = ∆uF B + ∆uF F
and then forming the control signal as
windup is avoided. This requires that the feedback control blocks have
inputs for feedforward signals.
Operational Aspects
Practically all controllers can be run in two modes: manual or automatic.
In manual mode the controller output is manipulated directly by the
operator, typically by pushing buttons that increase or decrease the con-
troller output. A controller may also operate in combination with other
controllers, such as in a cascade or ratio connection, or with nonlinear
elements, such as multipliers and selectors. This gives rise to more oper-
ational modes. The controllers also have parameters that can be adjusted
in operation. When there are changes of modes and parameters, it is es-
sential to avoid switching transients. The way the mode switchings and
the parameter changes are made depends on the structure chosen for the
controller.
282
6.7 Computer Implementation
Figure 6.21 Bumpless transfer in a PID controller with a special series imple-
mentation.
Figure 6.22 A PID controller where one integrator is used both to obtain integral
action in automatic mode and to sum the incremental commands in manual mode.
switch so that the signals are either chosen from the manual or the au-
tomatic increments. Since the switching only influences the increments
there will not be any large transients.
A similar mechanism can be used in the series, or interacting, imple-
mentation of a PID controller shown in Figure 6.22. In this case there
will be a switching transient if the output of the PD part is not zero at
the switching instant.
For controllers with parallel implementation, the integrator of the PID
controller can be used to add up the changes in manual mode. The con-
troller shown in Figure 6.22 is such a system. This system gives a smooth
transition between manual and automatic mode provided that the switch
is made when the output of the PD block is zero. If this is not the case,
there will be a switching transient.
It is also possible to use a separate integrator to add the incremental
changes from the manual control device. To avoid switching transients
in such a system, it is necessary to make sure that the integrator in the
PID controller is reset to a proper value when the controller is in manual
mode. Similarly, the integrator associated with manual control must be
reset to a proper value when the controller is in automatic mode. This can
be realized with the circuit shown in Figure 6.23. With this system the
switch between manual and automatic is smooth even if the control error
or its derivative is different from zero at the switching instant. When
the controller operates in manual mode, as is shown in Figure 6.23, the
feedback from the output v of the PID controller tracks the output u. With
efficient tracking the signal v will thus be close to u at all times. There
is a similar tracking mechanism that ensures that the integrator in the
manual control circuit tracks the controller output.
Figure 6.23 PID controller with parallel implementation that switches smoothly
between manual and automatic control.
283
Chapter 6. PID Control
Zt
xI = e(τ )dτ
K
I= xI
Ti
284
6.7 Computer Implementation
Computer Code
As an illustration, the following is a computer code for a PID algorithm.
The controller handles both anti-windup and bumpless transfer.
"Compute controller coefficients
bi=K*h/Ti "integral gain
ad=(2*Td-N*h)/(2*Td+N*h)
bd=2*K*N*Td/(2*Td+N*h) "derivative gain
a0=h/Tt
"Control algorithm
r=adin(ch1) "read setpoint from ch1
y=adin(ch2) "read process variable from ch2
P=K*(b*ysp-y) "compute proportional part
D=ad*D-bd*(y-yold) "update derivative part
v=P+I+D "compute temporary output
u=sat(v,ulow,uhigh) "simulate actuator saturation
daout(ch1) "set analog output ch1
I=I+bi*(ysp-y)+ao*(u-v) "update integral
yold=y "update old process output
The computation of the coefficients should be done only when the con-
troller parameters are changed. Precomputation of the coefficients ad, ao,
bd, and bi saves computer time in the main loop. The main program must
285
Chapter 6. PID Control
be called once every sampling period. The program has three states: yold,
I, and D. One state variable can be eliminated at the cost of a less readable
code. Notice that the code includes derivation of the process output only,
proportional action on part of the error only (b = 1), and anti-windup.
6.8 Summary
286
7
Loop Shaping
7.1 Introduction
This is the first Chapter that deals with design and we will therefore
start by some general aspects on design of engineering systems. Design
is complicated because there are many issues that have to be considered.
Much research and development has been devoted to development of de-
sign procedures and there have been numerous attempts to formalize the
design problem. It is useful to think about a design problem in terms of
specifications, trade-offs, limitations, design parameters. In additions it is
useful to be aware of the fact that because of the richness of design prob-
lem there are some properties that are captured explicitly by the design
method and other properties that must be investigated when the design
is completed.
The specifications is an attempt to express the requirements formally.
For control systems they typically include: attenuation of load distur-
bances, measurement noise, process uncertainty and command signal fol-
lowing. Specifications are typically expressed by quantities that capture
these features. For control systems the fundamental quantities are typi-
cally transfer functions or time responses and specifications are some rep-
resentative features of these functions. A design fundamentally involves
trade-offs. It is also extremely important to be aware of fundamental limi-
tations to avoid unrealistic specifications and impossible trade-offs. Some-
times it is attempted to capture the trade-offs in a single optimization
criterion, but such a criterion always includes parameters that have to be
chosen. It is often useful to make the trade-offs explicit and to have design
parameters that gives the designer control of them. A design method typi-
cally focuses on some aspects of a design problem but it is often difficult to
capture all aspects of a design problem formally. There will frequently be
287
Chapter 7. Loop Shaping
288
7.2 Compensation
7.2 Compensation
The idea of loop shaping is to find a controller so that the loop transfer
function has desired properties. Since the loop transfer function L = PC
is the product of the transfer functions of the process and the controller
it is easy to see how the loop transfer function is influenced by the con-
troller. Compensation is typically done by successive modifications of the
loop transfer function starting with proportional control. If the desired
properties cannot be obtained by proportional control the controller is
modified by multiplying the controller transfer function by compensating
networks having simple transfer functions. The design is conveniently
visualized using bode diagrams. Since the Bode diagram is logarithmic
multiplication by a compensating network corresponds to additions in the
Bode diagram. We start by discussing a desired properties of a loop trans-
fer function, then we present some simple compensators and we end by a
few examples. In the following sections we will present a more systematic
treatment where we discuss how th different properties of the loop are
influenced by the loop transfer function.
ϕ m arctan (2 + ng c)
289
Chapter 7. Loop Shaping
2
10
hG (ω gc )h
0
10
log10 gm
−2
10
−2 −1 0 1 2
10 10 10 10 10
ω
−120
−140
−160 ϕm
arg G (ω )
100(s+0.316)
Figure 7.1 Bode plot of the loop transfer function L( s) = s2 (s+3.16)(s+20)
.
The slope must be larger for non-minimum phase systems typical val-
ues for such systems are in the range of -1 to -0.5. The design starts
by adjusting the gain to give the desired crossover frequency compen-
sators are then added to give the desired phase margin and the desired
gain at low frequencies. The gain at low frequency can be increased by so
called lag compensators that increase the gain at low frequencies. The lag
compensators also decrease the phase of the loop transfer function. This
does however not cause stability problems because it happens far below
the crossover frequency. The phase margin is improved by so called lead
compensators which adds phase lead. Lead compensation required high
gain which means that measurement noise may be amplified. There are
also other types of compensators that decrease or increase gain at specific
frequencies.
Lag Compensation
Lag compensation is used to increase the gain of the loop transfer func-
tion at low frequencies. This will improve attenuation of low frequency
disturbances and reduce the error in tracking low frequency signals. A
290
7.2 Compensation
2
10
hG (ω gc )h
10
1
0
10
−2 −1 0 1
10 10 10 10
−20
arg G (ω )
−40
−60
PSfrag replacements
−80
−100
−2 −1 0 1
10 10 10 10
ω
Figure 7.2 Bode plot of the transfer function G ( s) = s+sa+/aM of a lag compensator
with a = 1 and M = 10 (full lines) and M = ∞ (dashed lines) which is a PI
controller.
where M > 1. This transfer function has unit gain for high frequencies,
ω >> a, and the gain M at low frequencies ω >> a. If M goes to infinity
the compensator becomes a PI controller with the transfer function
s+a a
G (s) = = 1+
s s
This gain of this compensator goes to infinity as s goes to zero. A Bode plot
of the compensator is given in Figure 7.2 The phase of the compensator is
always negative, which explains the name lag compensator. We illustrate
the use of lag compensation by an example.
291
Chapter 7. Loop Shaping
4
10
hG (ω gc )h
10
0
10
−2
10
−3 −2 −1 0 1
10 10 10 10 10
ω
−100
arg G (ω )
−150
−250
−3 −2 −1 0 1
10 10 10 10 10
ω
0.7
Figure 7.3 Bode plot of the loop transfer function L( s) =
s(s + 1)(s + 2)
(dashed lines) and the compensated loop transfer function Lc ( s) =
ki (s + 0.076)
(full lines).
s(s + 0.0038)(s + 1)(s + 2)
Assume that it is required to have a closed loop system with the following
specifications:
• No steady state error for constant load disturbances.
• A settling time to 1% that is shorter than 15s.
• The error in tracking the ramp input r = v0t should be less than
0.15v0.
A controller with integral action is required to avoid steady state errors.
Introducing an integrating controller with gain ki the loop transfer func-
tion becomes
ki
L(s) =
s(s + 1)(s + 2)
Choosing the gain ki = 0.7 and evaluating the closed loop system we
find that Ms = 1.4, ω ms = 0.64, ϕ m = 62○, and ω g c = 0.38, gm = 8.2
and a ω pc = 1.4 which are all reasonable values. A Bode plot of the loop
transfer function is shown in Figure 7.3. With M s = 1.4 and ω ms = 0.64
292
7.2 Compensation
we estimate the real part of the dominant poles to be faster than -0.3
which implies that the settling time is around 15 s. For small s the loop
transfer function and the sensitivity function are
0.7
L(s)
s
s
S(s) = 1.4s
0.7
This means that tracking error of the signal r = v0 t is 1.4v0 which is much
larger than the requirement of 0.04v0. To satisfy the specifications the low
frequency gain must be raised by a factor of 20. This can be achieved with
the lag compensator given by (7.1) with M = 20. The parameter a of the
compensator should be chosen so that the phase margin does not decrease
too much. Since the phase margin chosen originally was over 60 degrees
the choice a = ω g c/5 = 0.076 decreases the phase margin by about 11○
which can be acceptable. Adding the lag compensation the loop transfer
function becomes.
0.7(s + 0.076)
Lc (s) =
s(s + 0.0038)(s + 1)(s + 2)
The Bode plot of this loop transfer function is shown in Figure 7.2. An-
alyzing the system we find that M s = 1.5, ω ms = 0.55, ϕ m = 50○, and
ω g c = 0.34, gm = 7.6 and a ω pc = 1.3. For small s the loop transfer
function and the sensitivity function can be approximated by
14
L(s)
s
s
S(s) = 0.035s
0.7
The low frequency gain of the loop transfer function has thus been in-
creased substantially. The Bode plot in Figure 7.1 shows that the gain of
the compensated system is substantially higher for frequencies less than
0.02. To analyze the closed loop system obtained further we compute the
poles of the closed loop system and we find that the poles are -0.099,
−0.33 ± 0.36i and 2.24. The slow pole at -0.086 is close to the zero at
s = −0.076 so it has only a moderate influence on the step response. The
1% settling time is less than 5/0.33=15 s. Step and ramp responses of the
closed loop system are shown in Figure 7.4. Notice the that the compen-
sated system has a significantly better ramp response but also that it has
the inevitable overshoot. Compare with the discussion in Section 5.8.
293
Chapter 7. Loop Shaping
1.4
1.2
Step responses
1
0.8
0.6
0.4
0.2
0
0 5 10 15 20 25 30
t
30
Ramp responses
25
20
15
PSfrag replacements 10
0
0 5 10 15 20 25 30
t
Figure 7.4 Step responses and ramp responses for closed loop systems with error
0.7
feedback having loop transfer functions L( s) = (dashed lines)
s(s + 1)(s + 2)
ki (s + 0.076)
and Lc ( s) = (full lines).
s(s + 0.0038)(s + 1)(s + 2)
Lead Compensation
Lead compensation is used to improve the stability margin of a system or
to increase the crossover frequency. A lead compensator has the transfer
function
s+b
G (s) = (7.2)
s/ N + b
where N > 1. This transfer function has unit gain for low frequencies,
ω << b, and the gain N at high frequencies ω >> b. This compensator
has its largest gain, maxω h G (iω )h = N, at high frequencies. The largest
phase lead is given by
N−1
ϕ = max arg G (iω ) = arctan √
ω 2 N
√ √
is obtained for ω = ω m = a N. The gain at ω m is h G (iω m )h = N. The
achievable phase lead increases with N but it is less than 90○. Phase
leads 20○, 30○, 40○, 45○, 50○ and 60○ correspond to N = 2, 3, 4.6, 5.8 7.5
294
7.2 Compensation
2
10
hG (ω gc )h
10
1
0
10
−1 0 1 2
10 10 10 10
100
80
arg G (ω )
60
40
PSfrag replacements
20
0
−1 0 1 2
10 10 10 10
ω
s+b s
G (s) = = b+
b b
A Bode plot of the compensator is given in Figure 7.5 The phase of the
compensator is always positive, which explains the name lead compen-
sator. We illustrate the used of lead compensation by an example.
295
Chapter 7. Loop Shaping
4
10
2
10
hG (ω )h
0
10
−2
10
−2 −1 0 1 2
10 10 10 10 10
ω
0
−50
arg G (ω )
−100
−150
PSfrag replacements
−200
−250
−2 −1 0 1 2
10 10 10 10 10
ω
25
Figure 7.6 Bode plot of the loop transfer function L( s) = s2 +0.025s+0.25
(dashed
lines) and the compensated loop transfer function Lc ( s) = (s+15)(s25 (3s+5)
2 +0.0025s+0.25)
(full lines).
the frequency scale to krad/s instead of rad/s and the process transfer
function then becomes
0.25
P̄(s) =
s2 + 0.0025s + 0.25
To obtain the desired bandwidth the crossover frequency should be around
5 which means that the process gain has to be about 25. The Bode plot of
the loop transfer function with a gain of 25 is shown in Figure 7.6. The
figure shows that the phase margin is very small. To improve the phase
margin we introduce a lead compensator. To have reasonable damping we
choose N = 9 which gives a phase lead of 53○. Adjusting the gain so that
the the gain of the compensator is 1 at the crossover frequency ω g c = 5
we find that the compensator is
1 s + 5/3 3s + 5
G (s) = =
3 s/9 + 5/3 s + 15
and the compensated loop transfer function becomes’
25(3s + 5)
L(s) =
(s + 15)(s2 + 0.0025s + 0.25)
296
7.2 Compensation
2
10
hG (ω )h
10
1
0
10
−2 −1 0 1 2
10 10 10 10 10
100
50
arg G (ω )
PSfrag replacements
−50
−100
−2 −1 0 1 2
10 10 10 10 10
ω
(s+a)(s+b)
Figure 7.7 Bode plot of the transfer function G ( s) = (s+a/ M )(s/ N +b) of a lead-lag
compensator with a = b = 1, and M = N = 10 (full lines) and M = N = ∞ (dashed
lines), which is a PID controller.
Lead-Lag Compensation
Cascading a lead compensator with lag compensator gives a compensator
with the transfer function
(s + a)(s + b)
G (s) = (7.3)
(s + a/ M )(s/ N + b)
where M > 1 and N > 1. This compensator combines the features of lead
and lag compensators If N and M go to infinity the compensator becomes
a PID controller with the transfer function
(s + a)(s + b) a+b a s
G (s) = = + +
bs ab s b
297
Chapter 7. Loop Shaping
0
10
h G (ω g c)h
−1
10
−1 0 1
10 10 10
x
60
40
arg G (ω )
20
−40
−60
−1 0 1
10 10 10
ω x
Figure 7.8 Bode plot of the transfer function (7.4) of a notch compensator with
a = b = 1, ζ a = 0.05 and ζ b = 0.5 (full) and ζ a = 0.158 (dashed).
Notch Compensation
The compensators we have discussed so far have real poles and zeros. It
is sometimes advantageous to have compensators with complex poles and
zeros. One common situation is when it is desired to suppress transmis-
sion of signals at a particular frequency to avoid excitation of resonant
modes. This can be accomplished with the compensator
s2 + 2ζ a s + a2
G (s) = (7.4)
s2 + 2ζ b b + b2
298
7.2 Compensation
Table 7.1 Maximum gain of systems that give a specified phase lead. The table is
based on n first order systems with the transfer function (7.2).
gives high gain in a narrow frequency band which is very useful to sup-
press load disturbances of a specific frequency.
A compensator which gives a large phase lead will thus also have a large
gain.
299
Chapter 7. Loop Shaping
Im s
iR
Γ
ir γ
Re s
PROOF 7.1
The proof which is outlined in Bode’s book is a straightforward application
of residue calculus. The integral of the function ( G (s) − G∞ )/s along the
contour shown in Figure 7.9 is zero because the function is regular inside
the contour. The integral along the large semi circle goes to zero as R goes
to infinity. The integral along the small semi circle can be evaluated by
residue calculus. Hence
Z Z 0
ds dω
0 = ( G (s) − G∞ ) = ( A(ω ) − G∞ + iΦ(ω ))
s −∞ ω
Z ∞
dω
+ ( A(ω ) − G∞ + iΦ(ω )) + iπ ( G (0) − G∞ )
0 ω
It follows from ii) and iii) that G (0) = A(0) and G∞ = A(∞). Because of
iii) the integrals of A(ω ) cancel and the result follows from ii).
The theorem is called the phase-area theorem because the gain is pro-
portional to the area under the phase curve. It makes it possible to esti-
mate the gain required to obtain a certain phase lead. To keep the gain as
300
7.3 Bode’s Ideal Loop Transfer Function
b (u )
ϕo
u
c c c
low as possible it is useful to have phase lead only over a small frequency
range. To have robustness it is however required that the frequency inter-
val is not too short. A representative
R phase curve is shown in Figure 7.10.
For this curve we have φ (u)du = 2cϕ and the gain becomes
4c
N = e π ϕ = e2γ ϕ (7.8)
s n
L(s) = . (7.9)
ω gc
The Nyquist curve for this loop transfer function is simply a straight
line through the origin with arg L(iω ) = nπ /2, see Figure 7.11. Bode
called (7.9) the ideal cut-off characteristic. In the terminology of automatic
control we will call it Bode’s ideal loop transfer function.
One reason why Bode choose the loop transfer function given by Equa-
tion (7.9) is that it gives a closed-loop system where the phase margin is
insensitive to gain changes. Changes in the process gain will change the
crossover frequency but the phase margin is ϕ m = π (1 + n/2) for all values
301
Chapter 7. Loop Shaping
ω max
Figure 7.11 Nyquist curve for Bode’s ideal loop transfer function. Figure should
be redrawn, add circle, change slope to 30 degxs.
of the gain. The amplitude margin is also infinite. The slopes n = −4/3,
−1.5 and −5/3 correspond to phase margins of 60○, 45○ and 30○.
The transfer function given by Equation (7.9) is an irrational transfer
function for non-integer n. It can be approximated arbitrarily close by ra-
tional frequency functions. Bode realized that it was sufficient to approx-
imate L over a frequency range around the desired crossover frequency
ω g c. Assume for example that the gain of the process varies between kmin
and kmax and that it is desired to have a loop transfer function that is
close to (7.9) in the frequency range (ω min , ω max ). It follows from (7.9)
that
With n = −5/3 and a gain ratio of 100 we get a frequency ratio of about
16 and with n = −4/3 we get a frequency ratio of 32. To avoid having too
large a frequency range it is thus useful to have n as small as possible.
There is, however, a compromise because the phase margin decreases with
decreasing n and the system becomes unstable for n = −2.
The operational amplifier is a system that approximates the loop trans-
fer function (7.9) with n = 1. This amplifier has the useful property that
it will be stable under a very wide range of feedbacks.
302
7.3 Bode’s Ideal Loop Transfer Function
2
10
1
10
h L(iω )h
0
10
−1
10
−2
10
−1 0 1
10 10 10
−128
−130
−132
arg L(iω )
−134
−136
−140
−142
−1 0 1
10 10 10
ω
Figure 7.12 Bode diagram of the loop transfer function obtained by approximat-
ing the fractional controller (7.12) with the rational transfer function (7.13). The
fractional transfer function is shown in full lines and the approximation in dashed
lines. Redraw, line too heavy.
Fractional Systems
It follows from Equation (7.9) that the loop transfer function is not a
rational function. We illustrate this with an example.
k
P(s) = (7.10)
s(s + 1)
Assume that we would like to have a closed loop system that is insensitive
to gain variations with a phase margin of 45○. Bode’s ideal loop transfer
function that gives this phase margin is
1
L(s) = √ (7.11)
s s
303
Chapter 7. Loop Shaping
s+1 √ 1
C(s) = √ = s + √ (7.12)
s s
7.4 Robustness
The loop transfer function L is the central object in loop shaping design
and the crossover frequency ω g c is the primary design variable. To make
an effective design it is essential to understand how the specifications
are influenced by the loop transfer function. The robustness of a closed
loop system can be described by the sensitivity function and the comple-
mentary sensitivity function. In this section we will investigate how these
transfer functions relate to the loop transfer function and the crossover
frequency.
304
7.4 Robustness
ω pc ω pc
ω ms ω ms
PSfrag replacements PSfrag replacements
ω gc ω
ω gscc
ω sc
Figure 7.13 Nyquist curve of the loop transfer function with indications of gain
crossover frequency ω gc , phase crossover frequency ω pc , sensitivity crossover fre-
quency ω sc and the frequency where the sensitivity function has its maximum ω ms .
Process Variations
Robustness is captured by the sensitivity functions
1 T
S( L) = , T ( L) =
1+ L 1+ L
which are directly related to the loop transfer function. A good way to show
how the sensitivity functions depend on L is to plot the loci of constant
magnitude and constant phase of S and T in the L-plane.
The function S( L) is particularly simple. The loci for constant h S( L)h
are simply circles with center at -1 and the loci for constant phase of S( L)
are straight lines through the point -1. This is illustrated in Figure 7.14.
The to find the loci of constant magnitude and phase for the function
T ( L) we introduce L = x + i y. The condition that T is constant can be
written
L x + iy
hT h = = =M
1+ L 1 + x + iy
305
Chapter 7. Loop Shaping
ℑL ℑL
−1 ℜL ℜL
PSfrag replacements PSfrag replacements
−1
Figure 7.14 L-plane loci for constant gain and constant phase of the sensitivity
function S = 1/(1 + L).
Hence
x2 + y2 = M 2(1 + 2x + x2 + y2)
or
M 2 2 M2
2
x− + y =
1 − M2 (1 − M 2)2
This is a circle with center at M 2/(1 − M 2) and radius M /( M 2 − 1). If
you know complex functions you will realize this immediately because
the map S( L) is a Möbius transformations that maps circles to circles
and straight lines.
The loci for constant phase of T are circle through the points 0 and -1.
It follows from the geometric interpretation of division of complex numbers
that the circle corresponding to arg T = φ has center at −0.5 + 0.5i/ tan φ
and radius 0.5/h cos φ h. Figure 7.15 shows the loci in the L-plane of con-
stant magnitude and phase of the complementary sensitivity function. The
loci for constant phase and constant magnitude are orthogonal. They are
often drawn in the same figure, called Hall diagram. This diagram gives a
much better insight into permissible parameter variations that the crude
estimates given by Equation (5.14), because the figure shows directly the
effect of the loop transfer function on the closed loop transfer function T.
The transfer function T is also the closed loop transfer function from ref-
erence to the output for a system with error feedback. Figure 7.15 shows
that the widest margin for process variations is when the loop transfer
function is close to the line ℜ L = −0.5. The figure also shows that it is
essential not to have much variation in the gain for frequencies close to
ω pc.
306
7.5 Measurement Noise
ℑL
ℑL
ℜL
ℜL
PSfrag replacements PSfrag replacements
Figure 7.15 L-plane loci for constant gain (left) and constant phase (right) of the
complementary sensitivity function T = L/(1 + L).
N = log( L)
This means that instead of plotting the real and imaginary parts of L we
plot the logarithm of the magnitude and the argument, see vFigure 7.16.
It is also customary to let the argument be the horizontal axis. The loci
for constant magnitude and constant phase of T are orthogonal if natural
scales are used. In the practice that developed in control engineering it
became practice of choosing other scales which distorts the angles. The
Nichols chart and the Bode plots were very effective tools when computa-
tion was not widely available.
307
Chapter 7. Loop Shaping
Figure 7.16 The Nichols plot is a digram where the magnitude of L is plotted in
logarithmic scale as a function of the argument of L. The loci of constant magnitude
and phase of the complementary sensitivity function are also plotted in the diagram.
C C
Gun = = SC = (7.14)
1 + PC T
The loop transfer function PC typically has high gain for low frequencies,
and we have approximately
1
Gun ,
P
for low frequencies. The loop transfer function is typically small for high
frequencies and we have approximately
Gun C
308
7.5 Measurement Noise
log h C(iω )h
log K c √
log K
PSfrag replacements √
log K
ω gc log ω
Figure 7.17 Gain curve of the transfer function of a typical controller that pro-
vides phase lead.
Reasonable values of the M c are determined by the noise level and the
span of the control signal.
An Estimate of Mc
We will now investigate how M c depends on the crossover frequency ω g c.
Let the process transfer function be P(s). Since
1
h C(iω g c)h = (7.16)
h P(iω g c)h
it follows that
1
Mc ≥ (7.17)
h P(iω g c)h
The high frequency gain of the controller can, however, be significantly
larger when the controller has substantial phase advance as is illus-
trated in Figure Figure 7.17. The phase of the process transfer function
at the crossover frequency is arg P(iω g c ). The controller should thus have
a phase lead of
ϕ l = − arg P(iω g c ) − π + ϕ m. (7.18)
It follows from Equation (7.8) that this phase lead is associated with a
gain increase K = e2γ ϕ l . It is reasonable to assume that the controller
transfer function is symmetric around the neighborhood of the crossover
frequency, as shown in Figure
√ 7.17. The high frequency gain of the con-
troller is thus a factor of K larger than the gain at the crossover fre-
quency, which is given by Equation (7.16). We thus obtain the following
309
Chapter 7. Loop Shaping
1 √ 1
Mc = max(1, K ) = f mc(ω g c, ϕ m )
h P(iω g c)h h P(iω g c)h (7.19)
γ (− arg P( iω )−π +ϕ )
f mc(ω , ϕ ) = max(1, e )
1
Gxd
C
The loop transfer function typically has low gain at high frequencies. For
high frequencies we thus have the following approximation
Gxd P
where the gain of P typically decreases with frequency. We can thus expect
the transfer function G xd to have a maximum. We illustrate this by an
example.
310
7.6 Load Disturbances
1 b(s)
P(s) = =
(s + 1)2 a(s)
The controller
G g0 s 2 + g 1 s + g 2
C(s) = =
F s(s + f 1 )
where
f 1 = 4a − 2
g0 = 6a2 − 8a + 4
g1 = 5a3 − 4a + 2
g2 = 2a4
gives a closed loop system with the closed loop characteristic polynomial
s(s + 4a − 2)
Gxd =
(s2 + as + a2 )(s + a)(s + 2a)
1Gain curves for the transfer function are shown in Figure 7.18 for a = 1,
2 and 5. The corresponding values of the gain crossover frequencies are
ω g c = 0.7, 2.1 and 6.5.
The example shows that choice of crossover frequency has a major effect
on attenuation of load disturbances and that it is very advantageous to
have a high crossover frequency.
Estimating Mxd
In Example 7.4 the gain of the transfer function G xd has a maximum,
compare with Figure 7.18. We will now derive an approximation for the
largest gain
Mxd = max h Gxd (iω )h (7.20)
ω
311
Chapter 7. Loop Shaping
0
10
−1
10
h Gxd (iω )h
−2
10
−3
10
−4
10
−2 −1 0 1 2
10 10 10 10 10
4
10
3
10
h Gun (iω )h
2
10
PSfrag replacements 1
10
0
10
−2 −1 0 1 2
10 10 10 10 10
ω
Figure 7.18 Gain curves for the transfer functions G xd for crossover frequencies
ω gc = 0.7 (dotted), 2.1(dashed) and 6.5 (full). In the upper curve we also show
the gain curve for the process transfer function in thick full lines. The crossover
frequencies are also marked on this curve. Redraw figure only one curve!
P
h Gxd h = = h SPh < Msh Ph
1+ L
P hT h Mt
h Gxd h = = <
1+ L h Ch h Ch
Assuming for simplicity that the controller is designed with M s < M and
Mt < M we find that
1
h Gxd (iω )h < M min , h P(iω )h
ω h C(iω )h
It follows from the definition of gain crossover frequency that the terms
h P(iω )h and 1/h C(iω )h are equal for ω = ω g c. In the neighborhood of ω g c
we have ω n
h P(iω ) C(iω )h (7.21)
ω gc
312
7.6 Load Disturbances
The largest value of the right hand side occurs for ω > ω g c if the controller
gain decreases at the crossover frequency and for ω < ω g c if the gain
increases.
If the both controller and process gains decreases at the crossover
frequency we have
For controllers with phase advance the gain increases at the crossover
frequency. For a controller designed with a phase margin of ϕ m the phase
lead at the crossover frequency is
ϕ l = − arg P(iω g c ) − π + ϕ m.
It follows from Equation (7.8) that this phase lead is associated with a
gain increase K = e2γ ϕ l . If the controller transfer function is symmetric
around the neighborhood of the crossover frequency, as √ shown in Fig-
ure 7.17 the controller gain will drop with the factor of K below the
crossover frequency and we find that
1 √ 1 √
max h Gxd (iω )h = M max =M K = M K h P(iω g c)h
ω ω h C(iω )h h C(iω g c)h
Combining with the case when the controller gain was decreasing at the
crossover frequency we obtain the following estimate
√ 1
max h Gxd (iω )h = M h P(iω g c)h max(1, K) = f mc(ω g c, ϕ m )
h P(iω g c )h
(7.22)
where the function f (ω , ϕ ) is defined in (7.19). Check the following section
and remove unnecessary material
The typical situation is that the process gain decreases around ω g c. If
For processes where the magnitude of the loop transfer function decreases
313
Chapter 7. Loop Shaping
monotonically around the gain crossover frequency it follows that the first
term is the smallest for ω < ω g c and the second term is the smallest for
ω > ω g c. For processes where the controller gain increase monotonically
and the controller gain decreases monotonically with frequency we have
the following estimate.
For these processes the maximum occurs at frequencies that are higher
than the crossover frequency ω g c. For controllers with substantial phase
lead the maximum occurs instead for frequencies that are lower than ω g c.
For such frequencies the sensitivity function is close to one and we obtain
the estimate
1
hh Gxd hh < min . (7.24)
h C(iω g c)h
The phase lead required is
α = − arg P(iω g c) − π + ϕ m
Using Bode’s phase area formula the gain required to obtain this lead can
be estimated by
K = e2γ (− arg P(iω gc )π +ϕ m )
Assuming that the phase curve is locally symmetric around the crossover
frequency we get the following estimate
Combining this with the results for the case when the maximum occurs
for ω > ω g c we get
314
7.7 Limitations Caused by Non-Minimum Dynamics
0
10
−1
10
−2
10
y
−3
10
PSfrag replacements
h Gxd (iω )h −4
10
−2 −1 0 1 2
10 10 10 10 10
ω x
Figure 7.19 Gain curves for the transfer functions G xd for crossover frequencies
ω gc = 0.7, 2.1 and 6.5. The estimates of the maxima of hhG xd hh given by Equa-
tion (7.24) are marked by o:s.
where Pmp is the minimum phase part and Pnmp is the non-minimum
phase part. Let ∆ P(s) denote the uncertainty in the process transfer func-
tion. The factorization is normalized so that h Pnmp(iω )h = 1 and the sign
is chosen so that Pnmp has negative phase. The achievable bandwidth is
characterized by the gain crossover frequency ω g c.
315
Chapter 7. Loop Shaping
loop transfer function Pmp C is equal to Bode’s ideal loop transfer function
given by Equation (7.9). Equation (7.26) then becomes
π
arg Pmp (iω ) + arg C(iω ) = n (7.27)
2
where n is the slope of the loop transfer function at the crossover fre-
quency. If replace n by g c the Equation (7.27) is a good approximation for
other controllers, with no RHP poles and zeros, because the amplitude
curve is typically close to a straight line at the crossover frequency. It
follows from Bode’s relations (3.41) that the phase is ng cπ /2. If the con-
troller has poles or zeros in the right half plane its non-minimum phase
parts have to be included in Pnmp .
It follows from Equations (7.26) and (7.27) that the crossover fre-
quency satisfies the inequality
where
π
α = −π + ϕ m − ng c . (7.29)
2
This condition which we call the crossover frequency inequality gives the
limitations imposed by non-minimum phase factors. A straightforward
method to assess the crossover frequencies that can be obtained for a
given system is simply to plot the left hand side of Equation (7.28) and
determine when the inequality holds. The following example gives a sim-
ple rule of thumb.
316
7.7 Limitations Caused by Non-Minimum Dynamics
10
2 Gain curve
1
10
hG (iω )h 0
10
−1
10
−2
10
−1 0 1
10 10 10
ω
Phase curve
0
−100
−150
−200
−250
−300
−1 0 1
10 10 10
ω
Figure 7.20 Bode plot of process transfer function (full lines) and corresponding
minimum phase transfer function (dashed).
means that there are systems that cannot be controlled with sufficient
stability margins. If the process has an uncertainty ∆ P(iω g c ) the crossover
frequency inequality (7.28) becomes more stringent
z−s
Pnmp(s) = . (7.31)
z+s
317
Chapter 7. Loop Shaping
Notice that Pnmp should be chosen to have unit gain and negative phase.
We have
ω
arg Pnmp(iω ) = −2 arctan
z
Since the phase of Pnmp decreases with frequency the inequality (7.28)
gives the following bound on the crossover frequency.
ω gc α
≤ tan . (7.32)
z 2
Time Delays
The transfer function for such systems has an essential singularity at in-
finity. The non-minimum phase part of the transfer function of the process
is
Pnmp(s) = e−sL. (7.33)
We have
arg Pnmp (iω ) = −ω L.
It follows from the crossover frequency inequality, Equation (7.28),
that
π
ω g c L ≤ α = π − ϕ m + ng c = 2α . (7.34)
2
π
The simple rule of thumb (7.30) gives ω g c L ≤ 2 = 1.57. Approximating
the time delay by
1 − sL/2
e−sL =
1 + sL/2
we find that a time delay L corresponds to a RHP zero s = −2/ L, using
the results for systems with a right half plane zero gives instead ω g c L ≤ 2.
Time delays give an upper bound on the achievable bandwidth.
s+p
Pnmp (s) = (7.35)
s−p
318
7.7 Limitations Caused by Non-Minimum Dynamics
where p > 0. Notice that the transfer function is normalized so that Pnmp
has unit gain and negative phase. We have
p
arg Pnmp(iω ) = −2 arctan .
ω
It follows from the crossover frequency inequality, Equation (7.28), that
p
ω gc ≥ . (7.36)
tan α /2
( z − s)(s + p)
Pnmp(s) = . (7.37)
( z + s)(s − p)
ω p ω g c/ z + p/ω g c
arg Pnmp(iω ) = −2 arctan − 2 arctan = −2 arctan .
z ω 1 − p/ z
√
The right hand side has its maximum for ω = pz, hence
r
p
arg Pnmp (iω ) ≤ −4 arctan
z
z α π ϕm π
≥ tan2 = tan2 − + ng c . (7.38)
p 2 4 4 8
319
Chapter 7. Loop Shaping
Table 7.2 Achievable phase margin for for ϕ m = π /4 and ngc = −1/2 and different
zero-pole ratios z/ p.
320
7.7 Limitations Caused by Non-Minimum Dynamics
321
Chapter 7. Loop Shaping
pT ≤ 0.326 (7.42)
Other Criteria
The phase margin and the slope at crossover are quite crude measures.
The advantage in using them is that it is very easy to obtain the crossover
frequency inequality, (7.28). By carrying out detailed designs the results
obtained can be refined and other criteria can be considered. In this section
we have made designs that give M s = Mt = 2 and Ms = Mt = 1.4. The
results give the following design inequalities.
• A RHP zero z
(
ω gc 0.5 for Ms, Mt < 2
≤
z 0.2 for Ms, Mt < 1.4.
• A time delay T
(
0.7 for Ms, Mt < 2
ω g cT ≤
0.37 for Ms, Mt < 1.4.
• A RHP pole p
(
ω gc 2 for Ms, Mt < 2
≥
p 5 for Ms, Mt < 1.4.
322
7.8 Trade-offs
A time delay or a zero in the right half plane gives an upper bound of
the bandwidth that can be achieved. The bound decreases when the zero
z decreases and the time delay increases. A pole in the right half plane
gives a lower bound on the bandwidth. The bandwidth increases with
increasing p. For a pole zero pair there is a upper bound on the pole-zero
ratio.
The rules of thumb are very useful for making a preliminary estimate
of what can be achieved by control. They are particularly useful in pre-
liminary phases of projects based on codesign of processes and controllers
because they provide a quick indication of difficulties of control that can
be alleviated by a redesign of the process.
7.8 Trade-offs
323
Chapter 7. Loop Shaping
Examples
1
P(s) =
(s + 1)(0.1s + 1)(0.01s + 1)(0.001s + 1)
324
7.8 Trade-offs
0
10
hhG xd hh
−2
10
0 1 2
10 10 10
0
−100
arg P
−200 P
−300
0 1 2
10 10 10
PSfrag replacements 4
10
hhG un hh
2
10
0
10
0 1 2
10 10 10
ω gc
Figure 7.21 The trade-off plot for a process with transfer function P ( s) = 1/( s +
1)(0.1s + 1)(0.01s + 1)(0.001s + 1) which shows estimates of the maximum load
disturbance gain hh G xd hh, maximum noise gain hh Gun hh and arg P as functions of the
crossover frequency ω gc . The dashed curve in the middle plot also shows the estimate
of the maximum high frequency gain kh f of the controller. The points marked o show
the crossover frequencies achievable by I, P and D controllers.
1
P(s) =
(s + 1)4
325
Chapter 7. Loop Shaping
0
10
hhG xd hh
−2
10
−1 0 1
10 10 10
0
arg P , Pnmp
−100
−200 P
−300
−1 0 1
10 10 10
PSfrag replacements 4
10
hhG un hh
2
10
0
10
−1 0 1
10 10 10
ω gc
Figure 7.22 The trade-off plot for a process with transfer function P ( s) = 1/( s +
1)4 which shows maximum load disturbance gain hhG xd hh and maximum noise gain
hh Gun hh as functions of the crossover frequency ω gc . The dashed curve in the lower
plot also shows the estimate of the high frequency gain of the controller.
1
P(s) = e− s
10 ∗ s + 1
This is non-minimum phase because it has a time delay, but the dynamics
is lag dominated. The process has ω 90 = 0.3 and ω 180 = 1.6. The trade-off
plot is shown in Figure 7.23. The figure indicates that the crossover fre-
quency must be less than 1 rad/s, for that frequency we have hh G xd hh = 0.2
326
7.8 Trade-offs
0
10
hhG xd hh
−1
10
−2
10
−1 0 1
10 10 10
0
arg P , Pnmp
−100
−200 P
−300
−1 0 1
10 10 10
PSfrag replacements 10
4
hhG un hh
2
10
0
10
−1 0 1
10 10 10
ω gc
Figure 7.23 The trade-off plot for a process with transfer function P ( s) =
e−s /(10s + 1) which shows maximum load disturbance gain hhG xd hh, maximum noise
gain hh Gun hh and phase of process and non-minimum phase part as functions of the
crossover frequency ω gc . The dashed curve in the lower plot also shows the estimate
of the high frequency gain of the controller.
1
P(s) = e− s
0.1 ∗ s + 1
This is non-minimum phase because it has a time delay. The process has
ω 90 = 1.4 and ω 180 = 2.9. The trade-off plot is shown in Figure 7.24. The
figure indicates that the crossover frequency must be less than 1 rad/s,
for that frequency we have hh G xd hh = 2 and the controller gain is M c = 0.3.
327
Chapter 7. Loop Shaping
0
10
hhG xd hh
−1
10
−1 0 1
10 10 10
0
arg P , Pnmp
−100
−200 P
−300
−1 0 1
10 10 10
PSfrag replacements
2
10
hhG un hh
0
10
−1 0 1
10 10 10
ω gc
Figure 7.24 The trade-off plot for a process with transfer function P ( s) =
e−s /(0.1s + 1) which shows maximum load disturbance gain hhG xd hh, maximum
noise gain hh Gun hh and phase of process and non-minimum phase part as functions
of the crossover frequency ω gc . The dashed curve in the lower plot also shows the
estimate of the high frequency gain of the controller.
10(s − 10)
P(s) = e−0.0025s
(s − 1)(s + 20)
This system has poles and zeros in the right half plane and a time delay.
The trade-off plot is shown in Figure 7.24. The figure indicates that in
order to have a controller with reasonable robustness the crossover fre-
quency must selected between 1 and 10 rad/s. The figure also indicates
328
7.8 Trade-offs
1
10
hhG xd hh
10
−1
10
−2
10
−1 0 1 2
10 10 10 10
100
arg P , Pnmp
−100 P
−200
−1 0 1 2
10 10 10 10
PSfrag replacements
2
10
hhG un hh
0
10
−1 0 1 2
10 10 10 10
ω gc
Figure 7.25 The trade-off plot for a process with transfer function P ( s) =
100(s−20)
(s−0.5)(s+20)
e−0.005swhich shows maximum load disturbance gain hh G xd hh, maximum
noise gain hh Gun hh and phase of process and non-minimum phase part as functions
of the crossover frequency ω gc . The dashed curve in the lower plot also shows the
estimate of the high frequency gain of the controller.
Summary
Summarizing we have obtained approximate expressions for the largest
gain of the transfer functions from load disturbances to process output
and from measurement noise to controller output. These expressions can
be used to find a suitable value of the gain crossover frequency as a com-
promise between attenuation of load disturbances and injection of mea-
surement noise.
hh Gxd hh = M h P(iω g c)h
M
hh Gun hh = eγ (−π +ϕ m−arg P(iω gc ))
h P(iω g c)h
Notice that the right hand side of these expression depends on the process
transfer function, the design parameters M and ϕ m and the parameter γ .
329
Chapter 7. Loop Shaping
7.10 Insights
330
7.10 Insights
Dominant Poles
The behavior of a feedback system is typically dominated by a pair of com-
plex poles, which we call dominant poles. These poles are closely related
to the behavior of the loop transfer function at the crossover frequency. A
simple method for approximate determination of the dominant poles from
knowledge of the Nyquist curve of the loop transfer function will now be
given. Consider the loop transfer function L(s) as a mapping from the s-
plane to the L-plane. The map of the imaginary axis in the s-plane is the
Nyquist curve L(iω ), which is indicated in Figure 7.26. The closed-loop
poles are the roots of the characteristic equation
1 + L(s) = 0
The map of a straight vertical line through the dominant closed-loop poles
in the s-plane is thus a curve through the critical point L = −1 in the L-
plane. This curve is shown by a dashed line in Figure 7.26. Since the map
is conform, the straight line AP CP is mapped on the curve AC, which in-
tersects the Nyquist curve orthogonally. The triangle ABC is also mapped
conformally to AP BP CP . If ABC can be approximated by a triangle, we have
iω 2 − iω 1 1 + L(iω 2 )
σ = (1 + L(iω 2 )) (7.43)
L(iω 2 ) − L(iω 1 ) LP (iω 2 )
331
Chapter 7. Loop Shaping
1− s
P(s) =
s
s2 + 2as + a2
g(1 − s)
L(s) = =
s(s + f )
f − g = 2a
g = a2
f = a2 + 2a
g = a2
332
7.10 Insights
a2(1 − s)
L(s) =
s(s + 2a + a2 )
Hence
a2 (1 − iω ) a2 (1 − iω )(2a + a2 − iω )
L(iω ) = =
iω (2a + a2 + iω ) iω ((2a + a2 )2 + ω 2 )
a2(1 + 2a + a2 ) a2 (1 + 2a − ω 2 )
=− −i
(2a + a ) + ω
2 2 2 (2a + a2)2 + ω 2
√
The loop transfer function intersects the real axis for ω 180 = a2 + 2a. At
the intersection the loop transfer function has the value
a2 (1 + 2a + a2 ) a
L(iω 180 ) = − =−
(2a + a2 )2 + (2a + a2 ) a+2
The gain margin of the system is g m = 1 + 2/a which can be made arbi-
trarily close to one by making a sufficiently large.
The example shows that it is possible to have systems where the closed
loop poles are very well damped and still the closed loop system is ex-
tremely sensitive to parameters. The reason for this behavior is that the
system has a zero at s = 1 in the right half plane. It thus follows from
the results of Section 7.7 that the crossover frequency cannot be larger
than 0.5 to have a robust closed loop system. Choosing a large value of a
is equivalent to choosing a high crossover frequency. The example shows
clearly the necessity of being aware of the fundamental limitations caused
by non-minimum phase dynamics.
The result can also be explained by Equation (7.43) which implies
that the real part of the root σ can be large even if the distance 1 + L
from the critical point is small if the derivative LP is also small.
1 + L(s) = 0
333
Chapter 7. Loop Shaping
334
7.10 Insights
B
P=
A
and a controller with error feedback having the transfer function
G
C=
F
The complementary sensitivity function is
BG
T=
AF + B G
the gain of this transfer function is close to one for small frequencies.
The gain will increase at a low frequency zero and it will keep increasing
until a closed loop pole is encountered. To avoid having a large gain of
T and a poor sensitivity it is therefore necessary that slow process zeros
are matched with closed loop poles that are close to the zeros. Since there
is a natural tendency for the gain of T to decrease it is advantageous to
choose the closed loop poles slightly faster than the zeros. Slow process
zeros must be matched in the same way.
The reason why slow right half plane zeros impose severe limitations
is that they cannot be compensated by slow right hand poles, because this
will make the closed loop system unstable. Dass lässt tief Gucken
There are also restrictions imposed by fast process poles. To under-
stand this we will consider the process and the controller discussed above.
The sensitivity function is given by
AF
S=
AF + B G
The gain of the sensitivity function is 1 for high frequencies. Now consider
stable open loop poles that are faster than the gain crossover frequency. To
335
Chapter 7. Loop Shaping
AF + B G = ĀA f F + B̄ Bs G = A f Bs C̄
Ā F̄ + B̄ Ḡ = C̄
where the polynomial C has all its roots in the neighborhood of the gain
crossover frequency. The design is thus accomplished simply by cancelling
slow process zeros and fast process poles. The result is that the design
problem is simplified considerably. This is also one reason why successful
control design can be based on simple models.
We illustrate with an example.
(s + 0.1)(s + 0.2)
P(s) =
(s + 1)(s + 2)(s + 10)(s + 20)
336
7.11 Summary
7.11 Summary
7.12 References
337
8
Feedforward Design
8.1 Introduction
Y (s) P2 (1 − P1 G f f )
= = P2 (1 − P1 G f f ) S (8.1)
D (s) 1 + PC
338
8.2 Disturbance attenuation
PSfrag replacements
d
−F
r u y
Σ C Σ P1 Σ P2
−1
339
Chapter 8. Feedforward Design
Steam valve
Feed F F
water
L Drum
Oil Turbine
Air
Raiser Down comer
Applications
In many process control applications there are several processes in series.
In such cases it is often easy to measure disturbances and use feedfor-
ward. Typical applications of feedforward control are: drum-level control
in steam boilers, control of distillation columns and rolling mills. An ap-
plication of combined feedback and feedforward control follows.
340
8.3 System inverses
P−1 (s) = (1 + sT ) e sL
This system is not a causal dynamical system because the term e sL repre-
sents a prediction. The term (1 + sT ) requires and ideal derivative which
also is problematic as was discussed in Section 6.3. Implementation of
feedforward thus requires approximations.
341
Chapter 8. Feedforward Design
with feedback we can take that into account when finding approximate
inverses.
Let P† denote the approximate inverse of the transfer function P. A
common approximation in process control is to neglect all dynamics and
simply take the inverse of the static gain, i.e.
P† (s) = P(0)−1
1 + sT
P† (s) =
1 + sT / N
Feedforward can be used very effectively to improve the set point response
of the system. This was has been discussed in connection with systems
having two degrees of freedom in Section 5.3. Here we will go a little
deeper into the design of feedforward compensators.
342
8.4 Response to Reference Inputs
a
d n
v
r e u x y
F Σ C Σ P Σ
Controller −1
PSfrag replacements
Process
b
d n
r e u x y
F Σ CI Σ P Σ
v CR
−1
PSfrag replacements
P† M uf f
r
ym u y
M Σ C Σ P
−1
Figure 8.3 Block diagram of three system with two degrees of freedom.
343
Chapter 8. Feedforward Design
and the feedback controller has the transfer function C. The system in the
middle (b) is a slight modification where the controller transfer function
is written as
C = CI + CR
where CI is the term of the controller function that contains integral
action and CR is the rest of the controller transfer function. For the PID
controller
ki
C = k+ + kd s
s
we have
ki
CI =
s
CR = k + kd s
The block diagram at the bottom of the figure (c) is yet another config-
uration. In this system there is an explicit feedforward signal u f f that
generates an input to the process that gives the ideal response y m to a
reference input r. This input is generated by the transfer function M . The
feedback operates on the error y m − y. It can be shown that all configu-
rations are equivalent if all systems are linear and the transfer functions
are chosen properly.
There are however some differences between the systems from a prac-
tical point of view. Assume for example that the transfer function M that
gives the ideal response to reference inputs is such that the transfer func-
tion P−1 M is stable with the same number of poles and zeros. It then
follows that the scheme c) in Figure 8.3 will give the ideal response to
reference signals for all controllers C that stabilize the system. There is
thus a clean separation between feedback and feedforward. The scheme
a) in Figure 8.3 will give the ideal response if
P−1 M + CM = CF
1 + PC
F=M
PC
The transfer function F thus depends on C. This means that if the feed-
back controller C is changed it is necessary to also change the feedforward
F. Notice that the feedforward F must cancel zeros in PC that are not
zeros of M . The corresponding equation for scheme b) is
1 + PC
F=M
PCI
344
8.5 Summary
The transfer function CI = ki /s does not have any zeros. With this scheme
the feedforward F must cancel zeros in P that are not zeros of M .
Notice that in all cases it is necessary to have an invertible process
transfer function P. Approximate inverses must be used if this transfer
function has RHP zeros or time delays. In scheme c) the process inverse
appears in the combination P† M . This transfer function will not contain
derivatives if the pole excess of P is not larger than the pole excess of M .
For example, if
1
P(s) =
s(s + 1)
1
M (s) = 2
s +s+1
then
s(s + 1)
P−1 (s) M (s) =
s2 + s + 1
Notice also that in this case the steady state gain of the transfer function
P−1 (s) M (s) is zero.
8.5 Summary
345
9
State Feedback
9.1 Introduction
The state of a dynamical system is a collection of variables that permits
prediction of the future development of a system. It is therefore very natu-
ral to base control on the state. This will be explored in this chapter. It will
be assumed that the system to be controlled is described by a state model.
Furthermore it is assumed that the system has one control variable. The
technique which will be developed may be viewed as a prototype of an
analytical design method. The feedback control will be developed step by
step using one single idea, the positioning of closed loop poles in desired
locations.
The case when all the state variables are measured is first discussed
in Section 9.2. It is shown that if the system is reachable then it is always
possible to find a feedback so that the closed loop system has prescribed
poles. The controller does not have integral action. This can however be
provided by a hack using state augmentation.
In Section 9.3 we consider the problem of determining the states from
observations of inputs and outputs. Conditions for doing this are estab-
lished and practical ways to do this are also developed. In particular it is
shown that the state can be generated from a dynamical system driven by
the inputs and outputs of the process. Such a system is called an observer.
The observer can be constructed in such a way that its state approaches
the true states with dynamics having prescribed poles. It will also be
shown that the problem of finding an observer with prescribed dynamics
is mathematically equivalent to the problem of finding a state feedback.
In Section 9.4 it is shown that the results of Sections 9.2 and 9.3 can be
combined to give a controller based on measurements of the process output
only. The conditions required are simply that the system is reachable and
346
9.1 Introduction
347
PSfrag replacements
u ẋ R x y
B Σ C
Figure 9.1 Block diagram of the process described by the state model in Equa-
tion (9.1) .
dx
= Ax + Bu
dt (9.1)
y = Cx
A block diagram of the system is shown in Figure 9.1. The output is the
variable that we are interested in controlling. To begin with it is assumed
that all components of the state vector are measured. Since the state at
time t contains all information necessary to predict the future behavior
of the system, the most general time invariant control law is function of
the state, i.e.
u(t) = f ( x)
If the feedback is restricted to be a linear, it can be written as
u = − Lx + Lr r (9.2)
This control problem is called the pole assignment problem or the pole
placement problem.
348
9.2 State Feedback
Examples
We will start by considering a few examples that give insight into the
nature of the problem.
u = − l 1 x1 − l 2 x2 + L r r
p(s) = s2 + 2ζ ω 0 s + ω 20
349
Chapter 9. State Feedback
Comparing this with the characteristic polynomial of the closed loop sys-
tem we fin the following values of the We find that the feedback gains
should be chosen as
l1 = ω 20 , l2 = 2ζ ω 0
The closed loop system which is given by Equation (9.7) has the transfer
function
s
−1 −1 0
G yr (s) = 1 0
l1 s + l 2 Lr
Lr Lr
= 2 = 2
s + l2 s + l 1 s + 2ζ 0ω 0 s + ω 20
u = l1 (r − x1 ) − l2 x2 = ω 20 (r − x1 ) − 2ζ 0ω 0 x2
This polynomial has zeros at s = 0 and s = −l1 . One closed loop pole is
thus always equal to s = 0 and it is not possible to obtain an arbitrary
characteristic polynomial.
350
9.2 State Feedback
This example shows that the pole placement problem cannot be solved.
An analysis of the equation describing the system shows that the state x2
is not reachable. It is thus clear that some conditions on the system are
required. The reachable and observable canonical forms have the property
that the parameters of the system are the coefficients of the characteristic
equation. It is therefore natural to consider systems on these forms when
solving the pole placement problem. In the next example we investigate
the case when the system is in reachable canonical form.
Expanding the determinant by the last row we find that the following
recursive equation for the determinant.
Dn (s) = sDn−1(s) + an
A useful property of the system described by (9.8) is thus that the co-
efficients of the characteristic polynomial appear in the first row. Since
the all elements of the B-matrix except the first row are zero it follows
that the state feedback only changes the first row of the A-matrix. It is
351
Chapter 9. State Feedback
thus straight forward to see how the closed loop poles are changed by the
feedback. Introduce the control law
l̃1 = p1 − a1
l̃2 = p2 − a2
..
.
l̃n = pn − an
The system (9.10) has the following transfer function from reference to
output from
352
9.2 State Feedback
Notice that the system has the same zeros as the open loop system. To
have unit steady state gain the parameter L r should be chosen as
an + l̃n pn
Lr = = (9.12)
bn bn
z = Tx
Notice that the matrix W̃r is given by (3.64). The feedforward gain L r is
given by Equation (9.12).
The results obtained can be summarized as follows.
353
Chapter 9. State Feedback
dx
= Ax + Bu
dt
y = Cx
with one input and one output which has the transfer function
u = − Lx + Lr r
354
9.2 State Feedback
det(sI − A + B L)
355
Chapter 9. State Feedback
Notice that the program acker does not give the static gain L r , com-
pare (9.6), and that the notation is different from the one we use in the
book. This is easily dealt with by writing a new Matlab function sfb which
computes the gain matrices for the state feedback problem.
function [L Lr]=sfb(A,B,C,p)
% Compute gains L Lr for state feedback
% A, B and C are the matrices in the state model
% p is a vector of desired closed loop poles
L=acker(A,B,p);
Lr=1/(C*((-A+B*L)\B);
Integral Action
The controller (9.2) does not have integral action. The correct steady state
response to reference values was obtained by a proper choice of the gain
Lr , i.e. a calibrated procedure. Compare with (9.6). This means that the
controller is not useful practically. One way to ensure that the output will
equal the reference in steady state is enforce integral action. One way to
do this is to introduce the integral of the error as an extra state, i.e.
Z
xn+1 = ( y − r)dt
dxn+1
= y − r = Cx − r
dt
356
9.2 State Feedback
This equation has the same form as the original system (9.1). The reach-
ability matrix of the augmented system is
B AB ... An B
Wr =
0 CB ... CA n−1 B
To find the conditions for Wr to be of full rank the matrix will be trans-
formed by making column operations. Let a k be the coefficients of the
characteristic polynomial of the matrix A. Multiplying the first column by
an , the second by a n−1 and the (n-1)th column by a1 and adding to the
last column the matrix Wr it follows from the Cayley-Hamilton theorem
that the transformed matrix becomes
B AB . . . 0
Wr =
0 CB . . . bn
where
bn = C( An−1 B + a1 An−2 B + . . . + a n−1 B ) (9.16)
Notice that bn can be identified with a coefficient of the transfer function
of the original system
b1 sn−1 + b2 sn−2 + . . . + bn
G (s) =
sn + a1 sn−1 + . . . + an
357
Chapter 9. State Feedback
If the reference is constant r0 and the closed loop system is stable it follows
that
y0 = Cx0 = r0
The steady state output is thus equal to the reference for all values of the
gain Lr . This is no surprise because the control law (9.17) can be written
as
Z t Z t
u(t) = − Lx(t)− L I ( y(τ )−r(τ ))dτ + L r r(t) = Lr r(t)+ L I e(τ )dτ − Lx(t)
0 0
(9.18)
and it clearly has integral action. Compare with Equation (2.4). This com-
parison also shows that the term − Lx(t) is a generalization of derivative
action.
Summary
It has been found that the control problem is simple if all states are
measured. The most general feedback is a static function from the state
space to space of controls. A particularly simple case is when the feedback
is restricted to be linear, because it can then be described as a matrix or a
vector in the case of systems with only one control variable. A method of
determining the feedback gain in such a way that the closed loop system
has prescribed poles has been given. This can always be done if the system
is reachable. A method of obtaining integral action was also introduced.
A comparison with the PID controller showed that state feedback can
be interpreted as a PID controller where the derivative is replaced by a
better prediction based on the state of the system.
9.3 Observers
In Section 9.2 it was shown that the pole it was possible to find a feedback
that gives desired closed loop poles provided that the system is reachable
and that all states were measured. It is highly unrealistic to assume that
all states are measured. In this section we will investigate how the state
can be estimated by using the mathematical model and a few measure-
ments. Before reading this section it is recommended to refresh the mate-
rial on observability in Section 3.7. In that section it was shown that the
state could be computed from the output if the the system is observable.
In this Section we will develop some practical methods for determining
the state of the system from inputs and outputs. It will be shown that
the computation of the states can be done by dynamical systems. Such
systems will be called observers.
358
9.3 Observers
dx
= Ax + Bu
dt (9.19)
y = Cx
where x is the state, u the input, and y the measured output. The problem
of determining the state of the system from its inputs and outputs will
be considered. It will be assumed that there is only one measured signal,
i.e. that the signal y is a scalar and that C is a vector.
y = Cx
y = Cx
dy
= Cdxdt = CAx + CBu
dt
d2 y dx du du
= CA + CB = CA2 x + CABu + CB
dt2 dt dt dt
..
.
dn−1 y n−1 n−2 n−3 du dn−2u
= CA x + CA Bu + CA B + . . . + CB
dtn−1 dt dtn−2
359
Chapter 9. State Feedback
Notice that the matrix on the left-hand side is the observability matrix
Wo . If the system is observable, the equation can be solved to give
u
y 0 0 ⋅⋅⋅
0
d y
du
dt
CB 0 0 ⋅⋅⋅
dt
−1 1
x = Wo
.
− W −
o
..
.
.
.
.
.
.
dn−1
n−2 n−3
n 1
CA B CA B ⋅ ⋅ ⋅ CB d n
n
− 2
2
dt − dt −
(9.20)
This is an exact expression for the state. The state is obtained by differ-
entiating inputs and outputs. Notice that it has been derived under the
assumption that there is no measurement noise. Differentiation can give
very large errors when there is measurement noise and the method is
therefore not very practical particularly when derivatives of high order
appear.
d x̂
= A x̂ + Bu (9.21)
dt
To find the properties of this estimate, introduce the estimation error
x̃ = x − x̂
d x̃
= A x̃
dt
If matrix A has all its eigenvalues in the left half plane, the error x̃ will
thus go to zero. Equation (9.21) is thus a dynamical system whose output
converges to the state of the system (9.19).
360
9.3 Observers
The observer given by (9.21) uses only the process input u, the mea-
sured signal does not appear in the equation. It must also be required that
the system is stable. We will therefore attempt to modify the observer so
that the output is used and that it will work for unstable systems. Con-
sider the following
d x̂
= A x̂ + Bu + K ( y − C x̂) (9.22)
dt
observer. This can be considered as a generalization of (9.21). Feedback
from the measured output is provided by adding the term K ( y − C x̂).
Notice that C x̂ = ŷ is the output that is predicted by the observer. To
investigate the observer (9.22), form the error
x̃ = x − x̂
d x̃
= ( A − K C) x̃
dt
If the matrix K can be chosen in such a way that the matrix A − K C has
eigenvalues with negative real parts, error x̃ will go to zero. The conver-
gence rate is determined by an appropriate selection of the eigenvalues.
The problem of determining the matrix K such that A − K C has pre-
scribed eigenvalues is very similar to the pole placement problem that
was solved in Section 3.7. In fact, if we observe that the eigenvalues of
the matrix and its transpose are the same, we find that could determine
K such that AT − C T K T has given eigenvalues. First we notice that the
problem can be solved if the matrix
C T AT C T . . . A(n−1)T C T
p(s) = sn + p1 sn−1 + . . . + pn
361
Chapter 9. State Feedback
It follows from Remark 9.1 of Theorem 9.1 that the solution is given by
K T = p1 − a1 p2 − a2 . . . pn − an W̃oT Wo−T
dx
= Ax + Bu
dt
y = Cx
362
9.3 Observers
p(s) = sn + p1 sn−1 + . . . + pn
REMARK 9.2
The dynamical system (9.22) is called an observer for (the states of the)
system (9.19) because it will generate an approximation of the states of
the system from its inputs and outputs.
REMARK 9.3
The theorem can be derived by transforming the system to observable
canonical form and solving the problem for a system in this form.
REMARK 9.4
Notice that we have given two observers, one based on pure differentiation
(9.20) and another described by the differential equation (9.22). There are
also other forms of observers.
363
PSfrag replacements
Chapter 9. State Feedback
y
Σ
u x̂˙ R x̂ − ŷ
B Σ −C
−1
A
Figure 9.2 Block diagram of the observer. Notice that the observer contains a
copy of the process.
Duality
Notice the similarity between the problems of finding a state feedback and
finding the observer. The key is that both of these problems are equivalent
to the same algebraic problem. In pole placement it is attempted to find
L so that A − B L has given eigenvalues. For the observer design it is
instead attempted to find K so that A − K C has given eigenvalues. The
following equivalence can be established between the problems
A 1 AT
B 1 CT
L 1 KT
Wr 1 WoT
The similarity between design of state feedback and observers also means
that the same computer code can be used for both problems. To avoid
mistakes it is however convenient to have a special code for computing
the observer gain. This can be done by the following Matlab program.
function K=obs(A,C,p)
% Compute observer gain K
% A and C are the matrices in the state model
% p is a vector of desired observer poles
K=acker(A’,B’,p)’;
364
9.3 Observers
k1 = p1 = 2ζ ω
k2 = p2 = ω 2
365
Chapter 9. State Feedback
In this section we will consider the same system as in the previous sec-
tions, i.e. the nth order system described by
dx
= Ax + Bu
dt (9.24)
y = Cx
u = − Lx + Lr r
for the case that all states could be measured and in Section 9.3 we have
presented developed an observer that can generate estimates of the state
x̂ based on inputs and outputs. In this section we will combine the ideas
of these sections to find an feedback which gives desired closed loop poles
for systems where only outputs are available for feedback.
If all states are not measurable, it seems reasonable to try the feedback
u = − L x̂ + Lr r (9.25)
x̃ = x − x̂ (9.27)
366
9.4 Output Feedback
Since the matrix on the right-hand side is block diagonal, we find that
the characteristic polynomial of the closed loop system is
This polynomial is a product of two terms, where the first is the charac-
teristic polynomial of the closed loop system obtained with state feedback
and the other is the characteristic polynomial of the observer error. The
feedback (9.25) that was motivated heuristically thus provides a very
neat solution to the pole placement problem. The result is summarized as
follows.
dx
= Ax + Bu
dt
y = Cx
u = − L x̂ + Lr r
d x̂
= A x̂ + Bu + K ( y − C x̂)
dt
REMARK 9.5
Notice that the characteristic polynomial is of order 2n and that it can
naturally be separated into two factors, one det (sI − A + B L) associated
with the state feedback and the other det (sI − A + K C) with the ob-
server.
367
Chapter 9. State Feedback
PSfrag replacements r u ẋ R x y
Lr Σ B Σ C
A
Process
K Σ
− ŷ
R x̂
B Σ −C
A
Lr Observer
−L
Figure 9.3 Block diagram of a controller which combines state feedback with an
observer.
REMARK 9.6
The controller has a strong intuitive appeal. It can be thought of as com-
posed of two parts, one state feedback and one observer. The feedback
gain L can be computed as if all state variables can be measured.
368
9.4 Output Feedback
of the system. A complex model thus gives a complex controller. Also no-
tice that it follows from (9.29) that the transfer function of the controller
has the property that C(s) goes to zero at least as fast as s−1 for large
s. The approach thus results in controllers which have high frequency
roll-off. Compare with the discussion of frequency response in Section 5.5
where it was found that controllers with high frequency roll-off were de-
sirable in order to make systems less sensitive to model uncertainty at
high frequencies.
Notice that this equation has a triangular structure since the equation
for x̃ does not depend on x. Also notice that the reference signal does
not influence x̃. This makes a lot of sense because it would be highly
undesirable to have a system where reference inputs generate observer
errors.
369
Chapter 9. State Feedback
A state feedback was determined in Example 9.1 and and observer was
computed in Example 9.4. Assume that it is desired to have a closed loop
system with the characteristic polynomial
(s2 + 2ζ cω c s + ω 2c )(s2 + 2ζ oω o s + ω 2o )
l1 = ω 2c , k1 = 2ζ oω o
l2 = 2ζ cω c, k2 = ω 2o
It follows from (9.29) that the transfer function from y to u of the con-
troller is
s2 + 2ζ 2ω 2 s + ω 2
s2 + 2ζ 1ω 1 s + ω 21
It can be shown that the observation about the association of poles to state
feedback and observers is true in general and we can draw the following
important conclusion: Although it is convenient to split the design prob-
lem into two parts, design of a state feedback and design of an observer,
370
9.5 Comparison with PID Control
the controller transfer function is uniquely given by all the specified closed
loop poles. It does not matter if a pole is allocated by the state feedback
or the observer, all assignments will give the same transfer function from
y to u! Transfer functions between other signal pairs do however depend
on which poles are associated with the observer and the state feedback. If
reference values are introduced as described by (9.25) the transfer func-
tion from reference r to output y is uniquely determined by the poles
assigned to the state feedback.
Z t
dy
u = k(br − y) + ki (r(τ ) − y(τ ))dτ − Td
o dt
The differences between this controller and a PID controller are that there
are more terms and there is no integral action in (9.30). The estimates x̂i
in (9.30) are filtered through the estimator.
We can thus conclude that a PD controller can be interpreted as a
controller with feedback from two states where the first state is the output
and the second state is an estimate of the derivative of the output. We
can also conclude that the controller based on feedback from estimated
states lacks integral action.
371
Chapter 9. State Feedback
A Calibrated System
The controller based on state feedback achieves the correct steady state
response to reference signals by careful calibration of the gain L r and
that it lacks the nice property of integral control. It is then natural to
ask why the the beautiful theory of state feedback and observes does not
automatically give controllers with integral action. This is a consequence
of the assumptions made when deriving the analytical design method
which we will now investigate.
When using an analytical design method, we postulate criteria and
specifications, and the controller is then a consequence of the assumptions.
In this case the problem is the model (9.1). This model assumes implicitly
that the system is perfectly calibrated in the sense that the output is zero
when the input is zero. In practice it is very difficult to obtain such a
model. Consider, for example, a process control problem where the output
is temperature and the control variable is a large rusty valve. The model
(9.1) then implies that we know exactly how to position the valve to get
a specified outlet temperature–indeed, a highly unrealistic assumption.
Having understood the difficulty it is not too difficult to change the
model. By modifying the model to
dx
= Ax + B (u + v)
dt (9.31)
y = Cx
where v is an unknown constant we can can capture the idea that the
model is no longer perfectly calibrated. This model is called a model with
an input disturbance. Another possibility is to use the model
dx
= Ax + Bu
dt
y = Cx + v
372
9.5 Comparison with PID Control
373
Chapter 9. State Feedback
The first n rows of this matrix are linearly independent if the original
system is observable. Let a1 , a2 , . . . , an be the coefficients of the character-
istic polynomial of A. To find out if the last row is linearly independent,
we multiply the rows of the matrix with a n , an−1 , an−2, . . . , a1 and add to
the last row. It follows from Cayley-Hamilton Theorem that
C 0
CA CB
2
CA CAB
W̄o =
.
.
.
n−1 n−2
CA CA B
0 bn
where
The last row is linearly independent of the first n rows if the parameter bn
is different from zero. We can thus conclude that the state can be observed
if the original system is observable and if parameter bn is different from
zero. Notice that the condition for observability of the input disturbance is
the same as the condition (9.16) for reachability of the augmented system
used to introduce integral action in Section 9.2.
u = − Lx x̂ − v̂ + Lr r
d x̂ A − B Lx − K x C 0 x̂ K x B
=
+
y+
r
dt v̂ − Kv C 0 v̂ Kv 0
374
9.5 Comparison with PID Control
Notice that the system matrix of this equation has zero as an eigenvalue,
which implies that the controller has integral action. The transfer function
of the controller is given by
sI − A + B L + K C 0 −1 K
x x x
C(s) = Lx 1
Kv s Kv
1 1
= K v + ( Lx − K v C)(sI − A + B Lx + K x C)−1 K x
s s
The controller has integral action and the integral gain is
K i = K v 1 + C( A − B Lx − K x C)−1 K x
375
Chapter 9. State Feedback
s3 + k1 s2 + k2 s + k3
(s + αω )(s2 + 2ζ ω s + ω 2 )
gives the following expressions for the components of the observer gain K
k1 = (α + 2ζ )ω
k2 = (1 + 2αζ )ω 2
k3 = αω 3
s2( k1 l1 + k2 l2 + k3 ) + s( k2 l1 + k3 l2 ) + l1 k3
C(s) = (9.36)
s(s2 + s( k1 + l2 ) + k1 l2 + k2 + l1 )
The transfer function has two zeros and three poles, where one pole is at
the origin.
376
9.6 Disturbance Models
PSfrag replacements
Pulse Pulse Step Ramp Sinusoid
Examples of Disturbances
A suitable characterization of disturbances will first be given. To do this
we will first give some simple example of disturbances and their mathe-
matical models.
Some prototype disturbances like pulses, steps, ramps and sinusoids,
see Figure 9.4 , have traditionally been used to model disturbances. These
disturbances can all be described by simple dynamical systems as is il-
lustrated by the following examples.
377
Chapter 9. State Feedback
d2 v
= −ω 2 a sin ω t = −ω 2 v
dt2
dξ
= A vξ
dt (9.37)
v = Cvξ
378
9.6 Disturbance Models
This system is in standard form. Assuming that all states are measurable
it is natural to use the feedback
u = − Lx − v + L r r
Notice that the disturbance state is not reachable. This is natural because
the disturbance model (9.37) is not influenced by the control signal. In
spite of this the effect of the disturbances on the process can be elimi-
nated by proper choice of the gain Lv . Combining the control law with the
feedback law we obtain the following closed loop system
dx
= Ax + B (u + v) = ( A − B L) x + B (1 − Lv )v + Lr r
dt
u = − L x̂ − v̂ + Lr r
d x̂ A B x̂ B Kx (9.39)
=
+ u+
( y − C x̂)
dt v̂ 0 Av v̂ 0 Kv
379
Chapter 9. State Feedback
ŷ
uff
r Model and
Feedforward
xm State ufb u y
Generator Σ Σ Process
Feedback
- x̂
Observer
Figure 9.5 Block diagram of a controller based on a structure with two degrees
of freedom. The controller consists of a command signal generator, state feedback
and an observer.
380
9.7 Reference Signals
Details
Having described the key idea we will now consider the details. Let the
process be described by (9.1) and let P(s) denote the process transfer
function, Furthermore let M (s) denote the transfer function defining the
ideal response to the reference signal. The feedforward signal is then given
by
U f f (s) = M (s) P† (s) R(s) (9.40)
dxm
= Axm + Buff
dt (9.41)
ym = Cxm
u = uff + L( xm − x̂)
dxm
= Axm + Buff
dt
ym = Cxm
d x̂
= A x̂ + Bu + K ( y − C x̂ )
dt
381
PSfrag replacements
uff
r Model and ym
Σ e ufb u y
Feedforward
− Observer L Σ Process
Generator
Controller
Figure 9.6 Block diagram of a controller based on a structure with two degrees
of freedom. The controller consists of a command signal generator, state feedback
and an observer which estimates the error e = xm − x̂, between the ideal states and
the estimates states.
382
9.7 Reference Signals
1
P(s) =
s2
The transfer function which generates the feedforward signal uff is given
by
s2ω 2m
G f f (s) = M (s) P−1(s) =
s2 + 2ζ mω m + ω 2m
s2ω 2m 2ζ mω m s + ω 2m
M (s) P−1(s) = = ω 2m 1 −
s2 + 2ζ mω m + ω 2m s2 + 2ζ mω m + ω 2m
dz1
= −2ζ mω m z1 − ω m z2 + ω m r
dt
dz2
= ω m z1
dt
y m = ω m z2
uff = ω 2m (r − 2ζ m z1 − z2)
Notice that the same dynamical system generates both the desired model
output ym and the feedforward signal uff . Also notice that the gain in-
creases as the square of ω m . Fast responses can thus be obtained but
large control signals are required.
Combining the feedforward generator with the determination of an
output feedback in Example 9.5 we can obtain a complete controller based
on command following, state feedback and an observer for the double
integrator. The controller is a dynamical system of fifth order which is
383
Chapter 9. State Feedback
u = uff + ufb
uff = ω 2m (r − 2ζ m z1 − z2)
ufb = l1 e1 + l2 e2 + e3
y m = ω m z2
dz1
= −2ζ mω m z1 + ω m z2 + ω m r
dt
dz2
= −ω m z1
dt
de1
= e2 + k1 ( ym − y − e1)
dt
de2
= e3 + u + k2 ( ym − y − e1)
dt
de3
= k3 ( y − e1)
dt
s2ω 2m
G f f (s) =
s2 + 2ζ mω m + ω 2m
ω 2m
M (s) =
s2 + 2ζ mω m + ω 2m
( k1 l1 + k2 l2 + k3 )s2 + ( k2 l1 + l2 )s + k3 l1
G f b (s) =
s(s2 + ( k1 + l1 )s + k1 l2 + k2 + l1 )
( k1 l1 + k2 l2 + k3 )s2 + ( k2 l1 + l2 )s + k3 l1
L(s) =
s3(s2 + ( k1 + l1 )s + k1 l2 + k2 + l1 )
384
9.8 An Example
ϕ1 ϕ2
I
Motor
ω1 ω2
J1 J2
9.8 An Example
The Process
Consider the system shown in Figure 9.7. It consists of a motor that drives
two wheels connected by a spring. The input signal is the motor current
I and the output is the angle of the second wheel ϕ 2 . It is assumed that
friction can be neglected. The torque constant of the motor is kI , the
moments of inertia are J1 and J2 and the damping in the spring is kd .
The system is a representative example of control of systems with
mechanical resonances. Such systems are common in many different con-
texts, industrial drive systems, robots with flexible arms, disk drives and
optical memories. If the requirements on the control system are modest
it is possible to use a PID controller but the achievable performance is
limited by the controller. The PID controller will work quite well as long
as the rotors move in essentially the same way. When the system is driven
in such a way that the angles ϕ 1 and ϕ 2 starts to deviate substantially
the PID controller is not working well and superior performance can be
obtained by using a more complex controller.
The equations of motion of the system are given by momentum bal-
ances for the rotors
dω 1
J1 = kI I + k(ϕ 2 − ϕ 1) + kd (ω 2 − ω 1 )
dt
dω 1
J2 = k(ϕ 1 − ϕ 2) + kd (ω 1 − ω 2 )
dt (9.44)
dϕ 1
= ω1
dt
dϕ 2
= ω2
dt
Choose the state variables as
x1 = ϕ 1 , x2 = −ϕ 2
x3 = ω 1 /ω 0 , x4 = ω 2 /ω 0
385
Chapter 9. State Feedback
p
where ω 0 = k( J1 + J2 )/( J1 J2 ), which is the undamped natural fre-
quency of the system when the control signal is zero. The state equations
of the system are
0 0 1 0 0 0
1 dx
0 0 0 1
0
0
=
x +
u +
d
ω 0 dt
−α 1 α 1 −β 1 β1
γ 1
0
α2 −α 2 β 2 −β 2 0 γ2
J2 J1 kd kd
α1 = , α2 = β1 = , β2 =
J1 + J 2 J1 + J 2 ω 0 J1 ω 0 J2
kI 1
γ1 = 2 , γ2 =
ω 0 J1 ω 0 J2
386
9.8 An Example
2
10
0
10
g
−2
10
−1 0 1
10 10 10
−150
−200
−250
p
−300
−350
−1 0 1
10 10 10
w
Figure 9.8 Bode plot for the open loop transfer function from current to angle of
the second rotor.
robarmdata
eig(sys.A)
ans =
-0.0500 + 0.9987i
-0.0500 - 0.9987i
0.0000 + 0.0000i
0.0000 - 0.0000i
The system thus has two poles at the origin corresponding to the rigid
body motion and two complex poles corresponding to the relative motion
of the wheels. The relative damping of the complex poles is ζ = 0.05 which
means that there is very little damping. The transfer function from motor
current to the angle of the second rotor is
0.045s + 0.45
P(s) =
s4 + 0.1s3 + s2
PD Control
For low frequencies process dynamics can be approximated by the transfer
387
Chapter 9. State Feedback
function
0.45
P(s)
s2
Such a system can conveniently be controlled by a a PD controller. Requir-
ing that the closed loop system should have the characteristic polynomial
s2 + 2ζ 1ω 1 s + ω 21
1 + sTd
C(s) = k (9.45)
(1 + sTd/ N )2
388
9.8 An Example
g11 g12
2
10
0
10
1
10
0
−1
10
10
−1
10
−2 −2
10 10
−2 −1 0 1 −2 −1 0 1
10 10 10 10 10 10 10 10
g21 g22
−1 0
10 10
−2 −1
10 10
−3 −2
10 10
−2 −1 0 1 −2 −1 0 1
10 10 10 10 10 10 10 10
Figure 9.9 Gain curve of the frequency responses of the gang of four for PD
control of the system. The controller is given by (9.45) with parameters ω 1 = 0.15,
ζ 1 = 0.707 and N = 10.
by the response of P(s) C(s)/(1 + P(s) C(s)) can be reduced by set point
weighting. The load disturbance response can be improved at low fre-
quencies by introducing integral action. The peak in the response to load
disturbances can be reduced a little by further tuning of the parameters.
Notice that the high frequency resonant mode is traceable in the graphs
although the amplitude is small. The mode becomes very visible if the
parameter ω 1 is increased towards 0.3.
389
Chapter 9. State Feedback
g11 g12
1.5 60
50
1 40
30
0.5 20
10
0 0
0 50 100 150 200 0 50 100 150 200
g21 g22
1 1
0.5 0.5
0 0
−0.5 −0.5
0 50 100 150 200 0 50 100 150 200
Figure 9.10 Time responses of the gang of four for PD control of the system. The
controller is given by (9.45) with parameters ω 1 = 0.15, ζ 1 = 0.707 and N = 10.
L=acker(A,B,P);Acl=A-B*L;
Lr=-1/(C*inv(Acl)*B);
disp(’Feedback gain L=’);disp(L)
disp(’Refererence gain Lr=’);disp(Lr)
%Check the results
n=size(A);disp(’Specified closed loop poles’)
disp(P);
disp(’Closed loop poles obtained’)
disp(eig(Acl)’);
%Compute Closed loop transfer function
[num,den]=ss2tf(Acl,B*Lr,C,0);Gcl=tf(num,den);
disp(’Closed loop transfer function:’)
Gcl
u = − l 1 x1 − l 2 x2 − l 3 x3 − l 4 x4 + L r x r
390
9.8 An Example
0.4s + 4
G (s) =
s4 + 5s3 + 10s2 + 20s + 4
Notice that when using Matlab a few small terms appear because of in-
accuracies in the computations.
To find the stiffness of the system we calculate the transfer function
from a a disturbance torque on the second rotor to the deflection of the
rotor. This can be done by the Matlab script
%Computation of impedance
%This script computes the transfer function
%from a disturbance on the second rotor to
%the angle of that rotor
%Compute state feedback
robarmsfb;
%Compute the transfer function
A=sys2.A;B=sys2.B;C=sys2.C;
[num,den]=ss2tf(Acl,B,C,0);Gcl=tf(num,den);
disp(’Closed loop impedance function:’)
Gcl
%Compute stiffness at second rotor
ks=den(5)/num(4);
disp(’Stiffness=’)
disp(ks)
Executing the script we find that the transfer function from a torque
on the second mass to deflection is
391
Chapter 9. State Feedback
2 3
Out2 Out3
x’ = Ax+Bu x’ = Ax+Bu 1
y = Cx+Du y = Cx+Du
Step Out1
State−Space1 State−Space
Step1
1
y
−1
−2
0 2 4 6 8 10 12 14 16 18 20
10
4
u
−2
0 2 4 6 8 10 12 14 16 18 20
t
Figure 9.12 Simulation of the closed loop system with state feedback. The angle
ϕ 1 of the first rotor is shown in dashed curves, and the motion of the second rotor
in solid lines.
392
9.8 An Example
With a controller having integral action the closed loop system is of fifth
order and it is necessary to specify five poles. They are chosen as -3, -2,
-1, and −1 ± i. The design is executed with the following Matlab script.
%Design of State feedback for robotarm
%Get process model
robarmdata;A=sys2.A;B=sys1.B;C=sys2.C;
393
Chapter 9. State Feedback
The transfer function from reference to angle of the second rotor for the
closed loop system is
0.4001s2 + 4.801s + 8
G (s) =
s5 + 7s4 + 20s3 + 30s2 + 24s + 8
The transfer function from torque on the second rotor to its displacement
can be computed using the following Matlab script.
%Computation of impedance
%This script computes the transfer function
%from a disturbance on the second rotor to
%the angle of that rotor for a state feedback
%with integral action
%Compute state feedback
robarmsfbi;
%Compute the transfer function
B=[sys2.B;0];C=[sys2.C 0];
[num,den]=ss2tf(Acl,B,C,0);Gcl=tf(num,den);
394
9.8 An Example
0
10
−2
g 10
−4
10
−3 −2 −1 0 1 2
10 10 10 10 10 10
w
Figure 9.13 Gain curves for the open loop system (dotted), the system with state
feedback (dashed) and the system with state feedback and integral action (full).
395
Chapter 9. State Feedback
1
y
−1
−2
0 2 4 6 8 10 12 14 16 18 20
10
4
u
−2
0 2 4 6 8 10 12 14 16 18 20
t
Figure 9.14 Simulation of the closed loop system with state feedback having in-
tegral action. The angle ϕ 1 of the first rotor is shown in dashed curves, and the
motion of the second rotor in solid lines.
r=1;d=2;ts=20;[t,x,y]=sim(’robarmblk’,[0 ts]);
subplot(2,1,1);
plot(t,y(:,1),’r--’,t,y(:,2),’b’,t,ones(size(t)),’r:’,’linew’,2);
ylabel(’y’)
axis([0 20 -2 4])
subplot(2,1,2);plot(t,y(:,6),t,zeros(size(t)),’r--’,’linew’,2);
ylabel(’u’)
axis([0 20 -2 10])
The result of the simulation is shown in Figure 9.14. Compare this with
the simulation of the controller without integral action in Figure 9.14. The
responses to reference signals are quite comparable but the responses to
a disturbance torque are different. The controller with integral action
gives zero steady state error when a torque is applied. The first rotor
does however have a substantial deviation. This is necessary to create a
torque to oppose the disturbance.
396
9.9 Summary
9.9 Summary
397
Index
398
9.9 Summary
399
Chapter 9. State Feedback
400
9.9 Summary
401