Control Engineering - 1
Control Engineering - 1
Control Engineering - 1
Martin Braae
Department of Electrical Engineering
University of Cape Town
Licensed to
Control & Instrumentation Laboratory,
Department of Electrical Engineering,
University of Cape Town.
Published by the author
Produced by
MBuct, ElecEng, UCT.
iv
Contents
Contents .........................................................................................................iii
1. Introduction ...................................................................................... 1
1.1. Classification of Control Systems ...........................................................2
1.1.1. Logic Controllers........................................................................2
1.1.2. Continuous Controllers...............................................................3
1.2. Open Loop Control Systems ...................................................................4
1.2.1. Block Diagram for Open Loop Systems .....................................5
1.2.2. A Disadvantage of the Open Loop Configuration ......................6
1.3. Closed Loop Control Systems.................................................................7
1.3.1. Block Diagram for Closed Loop Systems...................................7
1.3.2. Advantages of Closed Loop Configurations ...............................8
1.3.3. Examples of Closed Loop Systems.............................................10
1.4. Common Features in Feedback Control Systems ....................................13
1.5 Design Procedure .....................................................................................14
Preface
These notes were written for a first course in Control Engineering given to Electrical
Engineering undergraduates (and their enlightening contributions are gratefully
acknowledged). Thus the notes emphasise electro-mechanical systems and try to
show how aspects of Linear Control Theory can be applied to the design of
electronic circuits that control processes with poor or bad dynamics and ensure
acceptable final systems.
The material is presented in a conversational style in that equations and figures flow
within the text. Essential mathematics cannot be avoided but hopefully the powerful
results that it produces can inspire the effort needed to master it.
A computer program was written to produce design plots and time simulations that
appear in this text. It is used extensively by students on the course and can be
ordered if required. (Details are given at the back of this book.)
Martin Braae
Oakridge,
June 1994
1
Introduction
Many engineering problems are concerned with the application of large forces, or
high power, to achieve some pre-defined result.
For example, a power transistor regulates the flow of large electric currents, an
amplifier pulses high voltages in a radio transmitter, a petrol engine powers a car, a
jet engine propels an aircraft, a mechanical shovel shifts large masses of earth,
conveyor belts move tonnes of material through unit processes in an industrial plant,
electric motors turn drums that reel in cables that hoist lifts up and down shafts in
both buildings and mines, and so on. The list is virtually endless.
Most power sources have become increasingly sophisticated over the years, as have
the associated control systems. In fact, many modern control systems perform
complex tasks which a human is physically unable to carry out since these are too
fast, too slow, too numerous, too monotonous or too dangerous. For example,
modern aircraft include many control systems that compensate automatically for
changes in flight conditions, provide auto-pilot facilities and even land the plane in
bad weather. As a result its performance is consistent under virtually all expected
operating regimes.
Thus a major concern of engineers today is the design and development of automatic
control systems. Such systems are found everywhere −− from the largest chemical
2 Control Engineering -1
plant to the smallest electronic circuit. The methods needed for designing control
systems to achieve specified results form the subject of this text.
Other systems use LOGIC signals (i.e. two-valued variables that are either On or
Off). Such systems are controllable by logic chips in digital electronic circuits, relay
racks for smaller industrial applications and PROGRAMMABLE LOGIC
CONTROLLERS (PLCs or PCs) for large installations.
Example
Consider a very small industrial system in which water is forced into a storage tank
by a centrifugal pump that is driven by an electric motor. Two level sensors are
positioned within the tank to monitor water level. These level sensors control the
operation of a single electrical switch that closes when the level falls below a
minimum value and opens when the level rises above a maximum value. During
shut-downs for maintenance, the safety of plant personnel is ensured by a lockable
off-switch. Another remote off-switch is provided for emergency stops. Finally a
temperature sensitive switch prevents thermal damage to the equipment.
The functions of the four switches that control the operation of the electric motor are
summarised below.
Switch Operation
S1 Closes when level low and opens when level high
S2 Locked open during maintenance
S3 Opened for emergency stops
S4 Opens when temperature is high
Ch 1: Introduction 3
This equipment and all its switches are interconnected to form a typical logic control
system, shown in the following schematic.
If signals in a system are CONTINUOUS (e.g. vary over a range, like the industrial
standards of 4 to 20[mA] or 0 to 10[V]) then a continuous controller is needed.
Example
This is a typical open loop system. Speed changes are made by altering the field
voltage but there is no check on whether the desired speed is attained or not.
Many other examples of open loop control systems exist in everyday life. For
example:
Again the list of systems is endless, but all its entries have one thing in common,
namely, that one variable, often low-power, is used to set the value of another
variable, often high-power.
Ch 1: Introduction 5
All these diverse systems can be represented schematically as a block with an input
and an output. Thus the electric motor (which converts field voltage, Ef, to speed, Ω
) has the block diagram:
Examples of block diagrams for some common open loop systems are:
In general, industrial control systems are quite complex and typically consist of a
number of blocks that are interconnected to form a larger system. Alternatively,
details of a large system, given as one block, can be defined in a block diagram that
shows its component sub-systems.
It is assumed that the interconnection of blocks does NOT alter the characteristics of
the individual blocks in any way. Thus voltages in the system of resistors:
6 Control Engineering -1
since its characteristics are very likely to be altered by interconnection with similar
systems. (More information is necessary to turn it into a true block diagram.)
Open loop configurations are preferred whenever possible due to their simple
structure, lower component count and ease of design. Unfortunately there are
numerous practical instances in which open loop control is not adequate, because its
characteristics are changing, it is subjected to external disturbances or it is unstable.
For example, a common problem with open loop systems is that the relationship
between the output, like rotational speed Ω(t) of an MG set, and the input, like its
field voltage Ef(t), is neither fixed nor predictable. (Note that such uncertainty in an
open loop system is an indication that it should be controlled in closed loop.)
Thus for a particular field voltage the actual rotational speed of the motor can vary
considerably, depending on the prevailing torque or load conditions. This is
indicated in the above sketch by the open loop operating line. Notice that setting the
field voltage does not ensure firm control over the motor speed.
Thus, if the load increases and the motor-generator system slows down, then the
field voltage is increased to ensure that the motor speeds up and remains at the
desired speed. Similarly, should the motor speed up, then the field voltage is
reduced to maintain its speed.
The closed loop system for the motor-generator system is more complex than the
equivalent open loop system as can be seen from its block diagram:
8 Control Engineering -1
The components and configuration of the system are chosen so that the actual speed
of the generator is regulated automatically at the desired value. To achieve this
effect, three new elements have been added to the original system. These are:
Feedback or Closed Loop control ensures greater accuracy for the relationship
between the input and the output variable than is achieved by open loop control.
To illustrate this, consider the following two systems, one in open loop the other in
closed loop.
In both instances a field control voltage of 10[V] is applied to a motor system and
ensures a speed of 1000[rpm].
Now assume that a fault occurs on the motor, or that its load changes, so that the
speed halves in the open loop system. In the closed loop system, feedback provided
by the tacho-generator compensates automatically for the changes that have
occurred in the motor characteristics, thereby minimising its drop in speed.
Observe in particular that the control system moves the process input to ensure that
its output remains constant. Graphically the steady state operating line for the closed
loop configuration is vastly different to that of the open loop system. The reason is
very simply that the closed loop reacts to compensate for unexpected effects, such as
changes in load.
Thus closed loop control gives improved precision for the process being controlled,
at relatively small additional cost. This is a general property of closed loop
feedback control systems and accounts for their popularity in engineering.
10 Control Engineering -1
The output voltage is related accurately to the input voltage with little sensitivity to
changes in the amplifier gain, A. All the components of this closed loop system are
electronic. (This circuit is so well known that its feedback configuration may not be
obvious, and it is dealt with later.)
A common method of level control. In more elaborate systems the float can be
designed to control a power source, such as a pneumatic or hydraulic valve, that in
turn controls the main water valve
Control engineers of today owe a great debt to designers who wrestled with the
practical aspects of speed regulation in steam turbines, shown schematically as:
This system has great historic significance since it is the first application of feedback
control in the Western World, devised by James Watt in 1769.
When the steam turbine's speed of rotation increases the weights on the governor are
thrown outwards under centrifugal forces. This closes the steam valve and hence the
turbine slows down. Similarly as the speed of rotation slows, the valve is opened
and the turbine speeds up. In this way the governor regulates system speed. The
main components of this control loop are mechanical and consequently more costly
to adjust than simple electronic components used in modern speed regulators.
An on-board control computer compares the course set for the ship with that
measured via a navigational satellite. The error is then used to adjust the rudder
angle and keep the ship on course.
Here paper arrives from the production process at a varying rate. The function of the
control system is to roll the paper onto a wind-up reel without tearing or distorting
it. The controller does this by maintaining constant tension.
• a CONTROL LAW that uses the setpoint and the output variable to
compute the best input variable continuously.
Once again the input is usually low-power and the output high-power.
Control Engineering deals with techniques that are used to design control laws
which ensure optimal performance for such closed loop systems. The processes that
can be controlled are diverse but in many instances the sensors that measure the
outputs, the actuators that set the inputs and the devices that implement the control
laws are electronic, sometimes analog but more often digital. In effect the
14 Control Engineering -1
Thus Electrical Engineers today are often involved in the design of control systems
and may even take on the role of Control Engineers who use electronics for process
optimisation.
• study the process and determine its primary function in terms of its
continuous signals,
• classify its signals as outputs and inputs, and identify failings of the
present process in its open loop configuration,
For example, in the typical configuration for a feedback or closed loop control
system:
mathematical models for the Plant (relating OUTPUT to INPUT) and the Controller
(relating INPUT to ERROR) are combined to form a mathematical model for the
open loop system (relating OUTPUT to ERROR) and more significantly for the
closed loop system (relating OUTPUT to SETPOINT).
The mathematical model chosen for describing the system should ideally:
Here the field voltage, Ef, is the input variable, u(t), and the speed, Ω, is the output
variable, y(t). Variations in Ef produce variations in Ω and the two signals can be
monitored continuously on an oscilloscope or a digital computer. Typically traces
might be:
Assuming that the system does not change with time (i.e. is time-invariant) and that
all transients have decayed, then, whenever Ef follows the shape shown above the
speed will respond with the same shape as previously.
The energy in the input pulse is normalised by giving it unit area. This pulse input is
then the unit pulse function. If its duration in time, ε, becomes extremely small it is
the UNIT IMPULSE FUNCTION (A function that is mathematically convenient but
unrealisable in practice).
The response of a system to a unit impulse input is known as its UNIT IMPULSE
RESPONSE or WEIGHTING FUNCTION, g(t). Note that the system response, y(t),
and hence its weighting function, g(t), is exactly zero for all time, t, less than zero.
This is essential for CAUSALITY.
For linear systems, if the impulse function is scaled to alter its area then the impulse
response is scaled proportionally. Thus an impulse of area A produces a response
y(t) = A g(t)
Also for linear systems, if two impulses of area A1 and A2 are applied to the system
then its response is the sum of the two impulse response functions:
2.1.1. Linearity
Most real physical systems are NOT linear but many can be assumed to exhibit
linear behaviour for small deviations about some operating point. Since many
analytical techniques in control engineering are only valid for linear systems
LINEARITY is an important concept.
Thus the output from a linear system can be predicted from any input by:
Example
Given the impulse response, g(t), for a particular system, compute its response to the
input function:
and combined to yield the system response to the given input function.
Mathematically the system response y(t) to input u(t) is computed from the
summation
y(t ) = u(0) ε g( t) + u( ε ) ε g( t - ε ) + u(2 ε ) ε g( t - 2 ε ) +...
t/ε
= å u( kε ) g(t - kε ) ε
k=0
where it is assumed that u(t) is zero for all time t < 0. (This is arranged easily in
practice by suitable definition of the input function.)
In the limit as ε → 0 the response, y(t), defined by the above summation converges
to the CONVOLUTION INTEGRAL
t
ò
y(t) = u(τ ) g(t − τ ) dτ
0
For example, the response to a UNIT STEP INPUT, which is more readily
implemented in practice, could as easily form the basis of a system model.
The prediction of system responses to arbitrary input functions is obtained from the
convolution integral
Ch 2: System Modelling 21
t t
ò ò
y(t) = u(τ ) g(t − τ ) dτ = u(t − τ ) g(τ ) dτ
0 0
Consider two system blocks, with weighting function models g1 and g2, connected
in cascade:
The overall model relating y(t) to u(t) is given by the weighting function
t
ò
g(t) = g1(τ ) g 2 (t − τ ) dτ
0
In this situation a model relating the output, y(t), to the setpoint, r(t), defines the
behaviour of the closed loop system. Now since
t
ò
y(t) = [r(τ ) − y(τ )] g(t − τ ) dτ
0
the closed loop system, h(t), can only be modelled implicitly in terms of the
weighting function for the open loop system.
d dn d dm
b o y( t) + b1 y( t) + ... + b n n y( t) = a o u( t) + a1 u( t) + ... + a m m u(t )
dt dt dt dt
where ai and bi are real constants and m ≤ n for realisable causal systems.
Example
This equation yields a compact parametric model for the electronic system. The
model is linear and can be used to characterise systems connected in cascade,
parallel and closed loop configurations. A useful feature of most differential
equation models is that the coefficients of the differential equations (and hence the
system response) can be related directly to the individual physical components
within the system (R and C here).
2.3.1. Predictions
It is very difficult to predict the response of a plant to an arbitrary input using a plant
model based on differential equations. However numeric integration methods like
Runge-Kutta algorithms or Predictor-corrector methods easily evaluate differential
equations using a digital computer. These may assist in predictions but give little of
the insight which is so essential to engineering design.
d2 d d
b2 2
y( t) + y(t) = a 2 y1 ( t) ... (2.3)
dt dt dt
d2 d d
b1b 2 y( t) + {b1 + b2 } y( t) + y( t) = a 2 {b1 y1 ( t) + y1 (t)}
dt 2 dt dt
= a 2a1u( t)
This result gives the differential model for the combined system.
d
b1 y1 ( t) + y1 (t) = a1u( t) ... (2.4)
dt
d
b2 y2 ( t) + y 2 (t ) = a 2 u( t) ... (2.5)
dt
d2 d d
b1 2
y1 (t) + y1 (t) = a1 u(t) ... (2.7)
dt dt dt
d2 d d
b2 2
y 2 (t) + y 2 (t) = a 2 u(t) ... (2.8)
dt dt dt
d2 d d d
b1b 2 2
y( t) + b2 y1 ( t) + b1 y 2 ( t) = {a1b 2 + a 2 b1} u(t ) ... (2.9)
dt dt dt dt
d2 d d
b1b 2 y( t) + {b1 + b 2 } y(t) + y( t) = {a1b2 + a 2 b1} u( t) + {a1 + a 2 }u( t)
dt 2 dt dt
Assume that the plant is modelled by the simple first order differential equation
26 ControlEngineering - 1
d
b y( t) + y(t ) = au( t)
dt
= a{r ( t) − y( t)}
where y(s) is the Laplace transform of variable y(t) and y(0) is the initial condition
for y(t) when t=0.
In most control applications the variables used within a plant model can be chosen
skilfully so that the initial conditions for the variables and their time-derivatives are
zero. This simplifies the analysis considerably, since
dk
L[ y( t)] = sk y(s)
dt k
Thus
Differentiation Multiplication
d/dt is equivalent to by s
in time domain in s-domain
Example
Ch 2: System Modelling 27
In practice this is achieved easily as illustrated below. The resulting variables have
simple Laplace transforms and are known as DYNAMIC variables.
To derive a model for any process the input variable is first held at a constant value.
Once the output variable steadies, the input is stepped to another value in a so-called
STEP TEST. The dynamic output variable is then defined on the graph by axes <t>
and <y(t)> as shown:
The dynamic input variable is similarly defined. Note that the time scales, t, of the
two dynamic variables, y(t) and u(t), are synchronised and that all initial conditions
of both y(t) and u(t) are zero for t<0.
The differential equation for the system is then transformed to its Laplace form
b0 y(s) + b1sy(s) + b2 s2 y(s)+... +b nsn y(s) = a 0u(s) + a1su(s) + a 2s2 u(s) +... +a msm u(s)
28 ControlEngineering - 1
with ai and bi real constants and m ≤ n. By re-arranging this equation, the Laplace
transform for the plant output, y(s), can be expressed as an explicit function of its
input, u(s), so that
The ratio of the two polynomials in 's' that arises in this equation is the TRANSFER
FUNCTION model, g(s), for the plant. So the plant is modelled by
y(s) = g(s) u(s)
The transfer function, g(s), defines the relationship between the system output and
its input and can be obtained directly from the Weighting Function, g(t), defined
earlier, since
g(s) = L[g(t)]
The transfer function model is linear and can be used to compute overall transfer
function models for systems connected in cascade, parallel and closed loop.
= g 2 (s) y1 (s)
Here
y(s) = y1 (s) + y 2 (s)
Here
y(s) = q(s) u(s)
= q(s) {r(s) - y(s)}
And so
q (s)
y(s) = r (s)
1 + q(s)
= h(s) r(s)
30 ControlEngineering - 1
where h(s) is the transfer function model for the closed loop system. It relates the
plant output, y(s), to its setpoint, r(s).
Clearly, of the modelling options considered, transfer functions are the easiest and
most convenient analytical tools for dealing with inter-connected systems found in
control engineering loops. Analytical methods for designing control systems rely
extensively on transfer function models, g(s), and these are now studied in
considerable depth.
Additional model types, like Frequency Response Models, State Space Models,
Pulse Transfer Function Models and Discrete Models, exist and are dealt with
later as required.
3
Common Mathematical
Models
When confronted with the task of deriving a mathematical model for a large inter-
connected industrial system, it is often convenient to start by splitting the original
system into simpler component sub-systems. These are then modelled individually
and re-combined later, using block diagram algebra, to provide the required model
for the entire system.
3.1.1. Resistor
is another possibility that applies to the resistor block diagram. The particular circuit
that is under investigation dictates which model form is required.
3.1.2. Capacitor
for the given block diagram, or, depending on the analytical requirements,
1
v1 (s) − v2 (s) = i(s)
sC
3.1.3. Inductor
1
i(s) = {v1 (s) − v2 (s)}
sL
or
v1 (s) − v2 (s) = sL i(s)
An Operational Amplifier (or OpAmp) multiplies its input voltage, eA, by a very
large negative gain, -A, to give the OpAmp output voltage, eo. Its circuit diagram is
given below and shows the supply rails as well as the input and output voltages.
A transistor is a current amplifier. When properly biased, its collector current, iC, is
β times its base current, iB. These two currents combine to form the emitter current,
iE, and the block diagram for the transistor on its own is:
The diagram also indicates where a load current, iL, would fit in the loop.
The circuit below is in fact a feedback configuration as shown by its block diagram.
Individual blocks in the loop are defined as transfer functions, so it is easy to derive
overall models for the circuit. Assume that voltage v2 is the output and voltage v1 is
the input. The appropriate transfer function model is then
v2 (s) sL
=
v1 (s) R + sL
Another model is possible if the output is taken to be the current, i(s). Its transfer
function is
i(s) 1/ R
=
v1 (s) 1 + sL / R
3.2.1. Spring
Mathematical models for the linear and the rotational springs are
F(s) = k{x1 (s) − x 2 (s)} and T(s) = k{θ1 (s) − θ2 (s)}
respectively, where F(s) is the net compression force acting on the spring in the
direction of x1(s) and T(s) is the net torque in the direction of θ1(s).
36 Control Engineering - 1
3.2.2. Damper
where F(s) is the net compression force acting on the damper in the direction of
x1(s) and T(s) is the net torque in the direction of θ1(s).
3.2.3. Inertia
where F(s) is the net compression force acting on the mass in the direction of x(s)
and T(s) is the net torque in the direction of θ(s).
Common Mathematical Models 37
3.2.4. Levers
3.2.5. Gears
The mathematical relationships between variables in the gear system are based on
the gear ratio, n, defined by
d2
n=
d1
1
Speed ratio: θ2 (s) = θ1 (s)
n
In practice it is often convenient to map inertia, torque and rotational speed through
a set of gears. Thereby every variable is referred to one shaft, before analysis starts.
38 Control Engineering - 1
For constant armature current, the torque developed by the motor is proportional to
field current and can be approximated by
T(s) = K t i f (s) ... (3.1)
The torque produced by the magnetic fields is balanced against the mechanical
damping and inertia so that
Hence, combining transfer function models implied in equations (3.1), (3.2) and
(3.3), the position of the output shaft, θ(s), is related to the field voltage, ef(s), by
the overall transfer function
θ (s) θ (s) T(s) i f (s)
=
e f (s) T(s) i f (s) e f (s)
1 1
= 2
Kt
sB + s J R f + sL f
Common Mathematical Models 39
K t / (B R f )
=
s{1 + s(J / B)}{1 + s(L f / R f )}
K
=
s{1 + sTm }{1 + sTf }
In many instances the field time constant is much faster than the mechanical time
constant (Tf << Tm) and so this third-order transfer function model can be simplified
to an approximate second-order model
θ (s) K
=
e f (s) s{1 + sTm }
The equation for the armature circuit is derived from its voltages
ea (s) = R a ia (s) + sLa i a (s) + e m (s)
θ(s) a0
=
ea (s) s{1 + b1s + b2s2 }
Its constants are derived by a detailed analysis of the block diagram for the motor,
which comprises the component sub-systems:
Note that the position of a load (or braking) torque in the loop is indicated on the
diagram. Once again there is a mechanical time constant (J/B) and an electric time
constant (La/Ra) associated with the loop, but they are not easily separated as in the
case of a field controlled motor.
3.3.3. dc Generator
Hence the transfer function model between the field voltage and the load voltage is
e L ( s) i g Z L e g i f Kg Z L
= =
e f (s) eg i f e f (R f + sL f )(R g + Z L + sL g )
Equations for more complicated, interconnected systems like the following motor-
generator set:
are easily developed using the previous models. The transfer function for the motor-
generator system is computed from the transfer function relationship
θ( s) θ (s) eg (s)
=
e f (s) eg (s) e f (s)
that combines the previous model for a motor with that for a generator.
Voltage eA is kept near zero by the operational amplifier and current iA is very
small. The block diagram for the system is derived directly by inspection of the
original circuit:
By means of block diagram algebra (explained later) the overall system becomes:
Common Mathematical Models 43
The torque, T(t), delivered by the motor is a complex non-linear function of its
speed of rotation, Ω(t), and the current, i(t), flowing through the control field coils.
44 Control Engineering - 1
To derive a linear mathematical model for the dynamics of this motor, it must be
assumed that the motor is operating at a particular speed, Ωo, and draws a current of
io to overcome the friction and any load torque. Under these conditions the non-
linear function, N, can be linearized about an operating point, (Ωo, io), as follows
∂N ∂N
∆T(t) = ( Ω0 , i 0 ) ∆Ω (t) + ( Ω0 , i 0 ) ∆i(t)
∂Ω ∂i
The electro-mechanical torque, T(s), produced by the motor must overcome drag
(B), inertia (J) and load torques (TL) within the system. From a torque balance
T(s) = {B + sJ}Ω(s) + TL (s)
Finally the current flowing in the field control coil is a function of the applied
voltage
e(s) = Z i(s)
where Z is the coil impedance, e(s) is r.m.s. voltage level and i(s) is the r.m.s.
current. So for small deviations about some (unspecified) operating point
∆e(s) = Z ∆i(s) ... (3.6)
From Equations (3.5) and (3.6) it is possible to draw a block diagram for the AC
motor system:
These blocks can be combined, as shown later, to give an overall transfer function
for the electric motor system. (Note that the ∆ symbol is omitted.)
Here the processes that are involved can be extremely complex, fundamental laws
cannot usually be applied and a heuristic approach is adopted. By measuring the
response y(t) to some input u(t), chosen to be simple (e.g. a STEP Function) or rich
in higher frequency components (e.g. PSEUDO-RANDOM Codes), the transfer
function is computed from the Laplace transformation
46 Control Engineering - 1
L[ y( t)]
g(s) = (Not always easy)
L[ u( t)]
Example
Consider a crusher used in the mineral extraction industry to break large rocks into
smaller particles.
In the crusher control scheme, speed of the conveyor belt is used to regulate the
feedrate of rocks into the crushing chamber. This in turn influences the power drawn
by the motor that drives the crusher. Thus the block for the crusher system has
Conveyor Speed as input and Crusher Power as output. The relationship between
these two variables is highly non-linear and once again the mathematical model,
albeit heuristic, is a linearized approximation that is valid only for a limited region
about the operating point (at which the experiments were conducted).
Here the output from one block is the input to the next. The combined system block
is determined by eliminating the common variable.
The same input goes to both blocks while the two outputs are combined additively.
The input to the forward block, q(s), is the difference between the setpoint and the
output of q(s) fed back through the feedback element f(s).
Comparators are easily moved but beware of the sign changes that are needed.
Rule (f)
Rule (h)
It is possible to apply the rules of block diagram algebra to simplify this diagram in
order to determine its transfer functions, relating speed, Ω, to control voltage, e, and
to load torque, TL. (The steps taken in the analysis are not unique but the final result
is.)
50 Control Engineering - 1
To simplify the expressions the blocks are given transfer function symbols, g1, g2, g3
and g4.
Using block diagram algebra these blocks are combined in stages. The rules that are
invoked are indicated within rectangles on the diagrams. It is vital to note that only
the items within the rectangle need be considered while applying the rule. First a
cascade and a comparator sub-section are simplified to yield:
Substitution of transfer functions, (g1, g2, g3 and g4) for the sub-systems gives the
relationship between shaft position and control voltage as well as the effect of the
load torque.
Note
The requirements of the actual problem dictate the manipulations that need to be
carried out on the block diagram. For example, a feedback design might not need
load torque explicitly, and the last step above would be unnecessary. Similarly an
investigation to re-design block g3 would keep it separate and not lump it with any
other blocks.
When used in a flexible manner block diagrams and their associated algebra can
become invaluable tools in formulating and analysing complex engineering systems.
4
Dynamic Variables and
Laplace Transforms
Once a transfer function model has been derived for the particular system under
investigation, control engineers are interested in predicting its dynamic performance.
For example, the system might be unstable and thus virtually useless, or it might be
too oscillatory or too slow (or both) to perform its specified task. Experienced
designers quickly and easily obtain such information from a detailed investigation of
the transfer function model itself, without considering its input and output in detail.
These conclusions are then exploited in the design to produce a satisfactory final
system that is stable, damped and fast.
The aim is to be able to draw sensible conclusions about transfer function models
like
y(s) 2 + 4s
= g(s) = ... (4.1)
u(s) 1.01 + 0.2s + s2
simply by studying the model itself, without computing its output for a given input.
(The model could represent either an open loop or a closed loop system.)
and know immediately that the system has a steady state gain of 1.98 and that its
transients oscillate with a frequency of 1[rad/s] and decay away in about 40[s].
Ch 4: Dynamic Variables and Laplace Transforms 53
In the field of process control, variables are often moved suddenly from one value to
another. The equivalent dynamic variable, which results when all initial conditions
are removed, is a STEP function.
Mathematically a step of amplitude a is defined in the time and the Laplace domains
as
ò
a
f(s) = ae −st dt =
=0 elsewhere s
0
ò
a
f(s) = at e −st dt =
=0 elsewhere s2
0
The IMPULSE function is defined by the Dirac Delta Function which is zero
everywhere, infinite at zero and has finite area.
=∞ when t=0 ò
f(s) = δ (t) e−st dt
0
=0 elsewhere
=1
f(t) = e -at
for t≥0 f(s) =
1
s+a
=0 elsewhere
and and
f(t) = cos(ωt) for t≥0 f(s) =
s
s2 + ω 2
56 Control Engineering - 1
and and
− at
f(t) = e cos(ω t) for t≥0 s+a
f (s) =
(s + a)2 + ω 2
If a function is shifted in time then only the phase angle of its Laplace Transform is
altered. Mathematically the relevant equation is
d nf
where f (n) (0) = Dn f(0) = (0) . Note the recursive nature of these equations.
dt n
Ch 4: Dynamic Variables and Laplace Transforms 57
1 1 1
L [f (-1) (t)] = f (s) + f ( −1) (0) → f (s)
s s s
1 1 1 1
L [f (-2) (t)] = f (s) + f ( −1) (0) + f ( −2) (0) → f (s)
s2 s2 s s2
:
:
1 1 1 1
L [f (-n) (t)] = f (s) + f (-1) (0)+... + f − n (0) → f (s)
sn sn s sn
It is worthwhile glancing back now at sections 4.1.1 to 4.1.4 (pages 53 to 54) and
observing carefully how, for example, a unit step function with zero initial
conditions is derived by integration of a unit impulse function, while a ramp arises
by differentiating a parabolic function, cosines from sines, etc.
Clearly, mixing basic Laplace transforms with appropriate properties provides the
engineer with a powerful set of analytical tools.
where the theorem only holds for f(s) with poles in the left-half of the s-plane (i.e.
for stable functions).
ò
1
= y(s) est ds
2π j
C
Luckily, the inverse transform is linear so this integral conversion is rarely required.
Most practical problems expand the original transform into a set of simpler
transforms using partial fractions, followed by standard inversions from tabulated
results.
Example (Inversion)
3(s + 1)
y(s) = (which is not in the tables)
s(s + 3)
To find the corresponding time function, y(t), start by expanding y(s) in partial
fractions. This gives
1 2
y(s) = +
s s+3
2
Find the value that y(t) reaches when the process model has its input suddenly
s+3
changed by 6 units. Thus the input function is given by the Laplace transform
6
u(s) = (From tables)
s
60 Control Engineering - 1
In the previous case, how fast does y(t) change immediately after its input is
stepped?
d
L[ y(t)] = sy(s) (Assuming zero initial conditions)
dt
2 6
=s (Derivative rules)
s+3 s
Hence
d
y(0) = lim s[sy(s)] = lim s2 y(s)
dt s→∞ s→∞
12
= lim s = 12
s→∞ s+3
Clearly the gain of g(s) is the final value of its response to an input that is a unit step
function.
5
Steady State Behaviour
and System Type Number
The final values attained by variables in closed loop systems are vital to their
acceptance in practice. For example, a system designed to track a setpoint that is
held at a constant value may never reach a setpoint that is constantly changing,
unless its controller is properly specified.
In fact, exact tracking of setpoints that demand constant position, constant velocity
or constant acceleration in the process output variable, is an important criterion for
many controller designs.
To illustrate the problem of steady state error, consider a two-stand steel rolling mill
in which an automatic feedback system is designed to control the steel thickness,
Y(t), to a set value, R(t).
−10
= [µm/N] (Note negative gain)
1 + 6s
Note that the controlled system, h(s), has a gain close to unity and is 60 times faster
than the original plant, g(s), in open loop. (It is also more robust.)
In spite of these improvements, the control is not satisfactory since it suffers from a
steady state error when the setpoint is held at a constant value (or position). This is
shown by analysis of steady state conditions in the loop:
The transfer function that relates the error signal to the system setpoint is derived
from block diagram algebra to be
1
e(s) = r(s)
1 + q(s)
1 + 6s
= r(s)
61 + 6s
10
To roll steel 10 [mm] in thickness requires a constant step input r(s) =
[mm].
s
The final value for the error signal is computed simply from the Final Value
Theorem and is
Ch 5: Steady State Behaviour and System Type Number 63
1 + 6s 10 10
lim se(s) = lim s = [mm] = 160 [ µm]
s→ 0 s→ 0 1 + 60s s 61
Thus, although the closed loop system improves on the plant dynamics, there is a
steady discrepancy between the specified setpoint, R(t), and the actual steel
thickness, Y(t). This error is a linear function of the setpoint, as shown:
Such steady state error is undesirable and can be reduced by increasing the open
loop gain of q(s) at low frequencies (i.e. where s → 0).
An integration circuit would eliminate the error completely at steady state, giving
steel of the exact thickness required. The reason for this is simply that an integrator
in k(s) gives q(s) infinite gain at s = jω = j0 (i.e. at steady state). Intuitively, any non
zero error will change the process input, until the error becomes exactly zero.
introduces the required integration into the open loop transfer function. It gives the
open loop transfer function
60(1 + 10s) 1 6(1 + 10s)
q(s) = = [µm/µm]
10s(1 + 6s) s (1 + 6s)
For a thickness setpoint of 10[mm], the steady state error attains the final value
s + 6s2 10
lim se(s) = lim s [mm]
s→ 0 s→ 0 6 + 61s + 6s2 s
64 Control Engineering - 1
=0
This example illustrates how a steady state error can exist in a closed loop system
with acceptable dynamics, unless the design of k(s) ensures that the error is exactly
zero.
For constant velocity inputs the setpoint is a ramp. Steady state error in this case is
known as VELOCITY ERROR and is zero only if two integrators are included in
q(s). The system is then known as a TYPE 2 SYSTEM.
For constant acceleration inputs the setpoint is a parabolic function of time, the
steady state error is known as ACCELERATION ERROR and is zero only if q(s)
contains three integrators. The system is then a TYPE 3 SYSTEM.
Each of the three most important cases is now considered in some detail.
In the case of the rolling mill, the position error constant is computed from
60
K p = lim = 60 under P control
s→ 0 (1 + 6s)
and
Ch 5: Steady State Behaviour and System Type Number 65
60(1 + 10s)
K p = lim =∞ under PI control
s→ 0 10 s(1 + 6s)
The Position Error is computed directly from the position error constant
1 1 1 1
e p = lim s = lim =
s→ 0 1 + q(s) s s→ 0 1 + q(s) 1 + K p
Note that the results are derived for a unit step change of 1[mm] in the setpoint and
must be scaled for a 10[mm] change.
which is derived using the Final Value Theorem on e(s) when r(s)=1/s2 (a ramp).
K a = lim s2 q(s)
s→ 0
1
ea =
Ka
5.5. Summary
An open loop system of the form
where constants a0 and b0 are non zero, is known as a TYPE M SYSTEM since it
contains M integrators. The error at steady state for such systems is dependent on:
(1) the class of setpoint applied to the closed loop system and
The effects of type number on final error are tabulated for the most important cases.
Error
System Position, ep Velocity, ev Acceleration, ea
Type 0 Finite Infinite Infinite
Type 1 Zero Finite Infinite
Type 2 Zero Zero Finite
Type 3 Zero Zero Zero
Example 1
Thus the steady state position, velocity and acceleration constants are
Kp = 10 Kv = 0 Ka = 0
The response of the closed loop system to unit step and unit ramp changes to the
setpoint are:
Clearly the position, velocity and acceleration errors in this example are
1 1 1 1
ep = = ev = =∞ ea = =∞
1 + K p 11 Kv Ka
with a setpoint, r(s), that is a Step, then a Ramp, and finally a Parabola.
68 Control Engineering - 1
Example 2
Thus the steady state position, velocity and acceleration constants are
Kp = ∞ Kv = 10 Ka = 0
And its responses to step and ramp changes in the setpoint are:
It should be noted that the velocity error is a constant error in position between the
actual output, y(t), and its setpoint, r(t).
Once again the results can be verified from the error transfer function
e(s) 1 s
= =
r(s) 1 + q(s) 10 + s
Ch 5: Steady State Behaviour and System Type Number 69
The majority of all Phase Lock Loop (PLL) design problems are amenable to
transfer function analysis. In control terminology the typical phase lock loop:
The control block k(s) fits into the continuous signal that drives the VCO (voltage
controlled oscillator). The form taken by the transfer function fq(s) is typically
K k
Type 1 fq = k(s) =
s(s + a) s+a
K(s + a ) k(s + a)
Type 2 fq = k(s) =
s2 s
K(s + a )(s + b) k(s + a)(s + b)
Type 3 fq = k(s) =
s3 s2
Clearly k(s) provides the required type number, stabilises the loop and compensates
the system to give acceptable dynamic response times.
some reference to its ability to track setpoints of a particular type with either zero
error, or alternatively, some small predefined steady state error. By computing the
relevant error constant it is possible to convert this requirement to a minimum type
number and a minimum gain for the open loop system.
Example
To track a unit parabolic setpoint, r(t), with a steady state error of less than five
[units] error between the setpoint and output, y(t), implies that
ea ≤ 5 [units] when r(t) = 0.5 t2
Thus, from the theory of error constants, the original specification implies that
K a = lim s2 q(s) ≥ 0. 2
s→ 0
1 1
For Type 1 the setpoint r(s) = and the appropriate controller is k(s) =
s s
1 1
For Type 2 the setpoint r(s) = and the appropriate controller is k(s) =
s2 s2
1 1
For Type 3 the setpoint r(s) = 3
and the appropriate controller is k(s) =
s s3
Conclusion
Extrapolating from these results, it seems likely that the controller form
1
k(s) =
s+a
Speculation
The response, y(t), of a particular system to a given input, u(t), or a setpoint, r(t), is
readily determined by Convolution or Laplace methods. However it is generally
quicker and easier to predict its performance directly from the relevant transfer
functions. This requires an in-depth understanding of the s-plane and is essential for
the design of successful control schemes.
the input or setpoint, u(s), and output, y(s), are related by the Laplace equation
y(s) = g(s) u (s)
where the transfer function, g(s), the input, u(s), and the output, y(s), are given as
rational functions in the complex Laplace variable, s.
Specifically
N(s) N u (s) N y (s)
g(s) = u(s) = and y(s) =
D(s) D u (s) D y (s)
where N(s), D(s), Nu(s), Du(s), Ny(s) and Dy(s) are polynomials in s.
Expansion of y(s) into partial fractions requires evaluation of the roots of equation
Dy(s) = 0 ... (6.1)
in the s-plane.
Depending on the polynomial coefficients, these roots can be real, complex or both.
For roots that are simple (i.e. distinct) the partial fraction expansion is easy to
perform. An added complication arises when roots are repeated.
Example
Computation of the system response, output y(t), starts with the expansion of
Equation (6.2) into partial fractions. The four roots of Equation (6.1) for this
denominator are
r1 = 0 , r2 = −2 , r3 = −1 + 3 j and r4 = −1 − 3 j
The roots are simple. Two are real and two are complex. The latter form a complex
conjugate pair. Thus the partial fraction expansion of y(s) is
k1 k2 k3 k4
y(s) = + + +
s (2 + s) (1 + 3j + s) (1 − 3j + s)
where ki are constants, known as RESIDUALS and computed from the limits
18
k1 = lim sy (s) = lim = 0. 9
s→ 0 s→ 0 (2 + s)(1 + 3j + s)(1 − 3 j + s)
18
k 2 = lim (s + 2)y(s) = lim = −0 . 9
s→−2 s→−2 s(1 + 3j + s)(1 − 3 j + s)
18
k3 = lim (s + 1 + 3j)y(s) = lim = −0 . 3 j
s→−1− 3j s→−1− 3j s(2 + s)(1 − 3j + s)
18
k4 = lim (s + 1 − 3j)y(s) = lim = 0. 3 j
s→−1+ 3j s→−1+ 3j s(2 + s)(1 + 3j + s)
(The residuals from complex conjugate pairs are themselves complex conjugates.)
Hence the expanded Laplace transform for the output is
0. 9 −0. 9 −0.3 j 0.3 j
y(s) = + + +
s (2 + s) (1 + 3j + s) (1 − 3j + s)
= 0. 9 − 0. 9 e −2t − 0 . 6 e − t sin(3t)
Ch 6: Prediction of System Response 75
Note
(3) the real part of real and complex roots sets its transient decay rate,
(4) if the real part of any root is positive (>0) then its corresponding term
in the response increases exponentially with time and the overall
response of the system is unstable,
(5) the final response contains components that can be traced back to
either the transfer function g(s) or its input u(s).
Observations
(2) Each root, ri, must be zero or negative for y(t) to be a constant or a
decaying function of time respectively (i.e. ri≤0).
(3) If any root ri is positive (i.e. ri>0) then that term, and hence y(t),
increases without bound.
(4) The roots arising from the D(s) factor in Dy(s) are of greatest interest,
since the control engineer can design these to be in optimal positions,
while the roots of the Du(s) factor are caused by the input u(s) and are
usually unpredictable.
For real systems, complex roots of Dy(s)=0 ALWAYS occur in pairs that are
complex conjugates. In essence these roots are handled in the same way as real
roots. (The only difference is that the roots and their associated residuals are
complex numbers as opposed to real numbers.)
where root rc is the complex root and rc* is its complex conjugate. Then
η
kc k *c ki
y(s) = + +å
(s − rc ) (s − rc ) i=3 (s − ri )
*
The residuals kc and kc* are complex conjugate constants computed from
Ch 6: Prediction of System Response 77
Observation
(2) Its frequency of oscillation is given by the imaginary part, 'b', of the
complex root, rc.
(3) This oscillation decays at a rate determined by the real part, 'a', of the
complex root, rc, and 'a' must be zero or negative for a bounded
response in y(t) (i.e. a≤0).
The Residuals for the repeated root are computed by a series of differentiations and
limits:
78 Control Engineering - 1
1 dµ
k q − µ = lim [(s − r )q y(s)] for µ=0, 1, 2, ... ,q-2
s→ r µ! d s µ
k 3t 2 k q −1t q − 2 k q t q −1 rt η
y(t) = [k1 + k 2 t + +... + + ]e + å k i e ri t
2! (q − 2)! (q − 1)! i=q+1
Observation
(1) The repeated root must be negative only (i.e. r<0) for a bounded
response.
(2) All the other roots must be zero or negative to ensure a bounded
response in y(t), as before (i.e. ri≤0).
6.4. Summary
The time-domain response y(t) of a system where y(s) is defined by
N y (s) N(s) N u (s)
y(s) = =
D y (s) D(s) D u (s)
With the insight that the Laplace variable gives into the behaviour of differential
equations, it is possible to modify the positions of some roots of Dy(s), by altering
D(s), until the final system performs as required by its specifications.
Ch 6: Prediction of System Response 79
Consider the case where y(s) is given by a function containing a single real root, r,
1
y(s) =
s−r
The correspondence between the position of this root in the s-plane and the time
domain response of y(t) is illustrated by graphs showing both root position and
system impulse response.
Note that the response, y(t), is STABLE (i.e. bounded) when r≤0 and UNSTABLE
(i.e. unbounded) when r>0.
Note
The slope of the curve <ln(y)> versus <t> gives the value of 'r'.
The correspondence between positions of these roots in the s-plane and the time
domain response of y(t) is illustrated by graphs showing the root positions and
corresponding system impulse response.
Note that the response, y(t), is STABLE (i.e. bounded) when a≤0 and UNSTABLE
(i.e. unbounded) when a>0. Also the frequency at which the response oscillates is
given by the value of 'b' (often in [radians/second])
Notes
(1) The period of these oscillations gives 2π/b and hence 'b'.
Control engineers design the system transfer function, g(s), to meet specifications
set for the system, but have little or no control over the input function u(s). Thus the
polynomial D(s) associated with the system transfer function, g(s), is of greatest
interest in the analysis and design of control systems. Its importance is reflected in
its name, since D(s) is known as the CHARACTERISTIC FUNCTION of the
system.
82 Control Engineering - 1
The roots of the Characteristic Equation are known as THE POLES of the system
and form the basis for linear control engineering theory.
Roots of the equation Du(s)=0 on the other hand merely characterise the system
input, u(s), and have little bearing on system performance (i.e. stability and transient
response). These roots influence the choice of system type number and are important
in some applications but the designer has no direct control over them and usually
concentrates on D(s).
Each mode, gi(s), is defined by either a single pole or a single pair of complex
conjugate poles. Analysis of these modes gives invaluable insight into the
performance of the original system.
Example
Ch 6: Prediction of System Response 83
These two single-pole sub-systems are clearly stable and so g(s) itself must be
stable. Further deductions can be made, if required.
A
Its transfer function is g(s) =
1 + sT
When perturbed by an input the resulting transients in the system response decay
away within a time period determined by T. Specifically its responses to Step and
Impulse perturbations of the input u(t) are:
84 Control Engineering - 1
B
Step Response, u(s) = Impulse Response, u(s) = B
s
BA − t / T
y(t) = BA (1 − e − t / T ) y(t) = e
T
Practical Application
Another example of its use is in the specification of pole positions in the s-plane.
Should the system under consideration be required to settle to within the ±2% band
in less than 10 seconds, then this criterion translates to the equivalent requirement
that the system have a pole that lies to the left of s = −0 . 4 in the s-plane.
Ch 6: Prediction of System Response 85
Aω 2n
The transfer function is g(s) =
s + 2 ζω ns + ω 2n
2
The pole positions for a second order system with ζ<1 lie along the following lines
in the s-plane
Note
When the damping factor ζ=0.7071 the complex poles lie at 45° to the σ-axes and
the system gives a good response that is Fast with Some overshoot but No
oscillation. This is generally considered to be ideal for mechanical systems where
excessive oscillations lead to accelerated wear-and-tear or metal fatigue, but for
chemical systems a lower damping factor (ζ=0.3) can often be tolerated and gives a
faster response at the cost of more oscillation.
The response of a second order system (with A=1 and ζ<1) to Unit Step and
Impulse perturbations of the input u(t) are
1
Unit Step Response, u (s) = Unit Impulse Response, u (s) = 1
s
ωn σ t ω 2n σ t
y(t) = 1 − e sin ( ω d t + φ ) y(t) = e sin( ω d t)
ωd ωd
Practical Application
Once again this information can be used to estimate a model, g(s), for a plant that
responds to a step input in an oscillatory manner. Also the time domain
specifications of, say, no oscillation and a settling time of 12 [s] to the 5% band will
impose the restriction on the system poles that
ζ > 0.7071
a < −0. 25
Ch 6: Prediction of System Response 87
For any given design it is possible to define an area in the s-plane within which the
poles of the final system must lie in order to meet required specifications.
As an example, poles positioned within the shaded region above indicate a system
that is slightly under damped with ζ>0.7071 and has transients that decay to within
the ±2% band in less than 5 [s].
analysis were undertaken, the problem could very easily become extremely complex
and completely insoluble. Thus it is important to exercise some engineering
judgement when dealing with real-world problems. In control systems this usually
translates to knowing which sub-systems need to be modelled as dynamic elements
(with poles) and which can be approximated by constants. The following example,
though simple, illustrates this point.
The complicated relationship that exists between the position of the control rods and
the temperature of the reactor is approximated by the transfer function
y(s)
g(s) =
u(s)
5
=
(s + 10)(s + 0.1)(s2 + 0. 2s + 1)
(3) Pole No 2 is closer to the origin than pole pair No 3 and its
characteristic response will dominate the system response to step,
ramp and parabolic inputs functions.
Such observations and predictions are easily made by inspection of system pole
positions in the s-plane. The conclusions are confirmed by comparing step responses
of the original system
5
g(s) =
(s + 10)(s + 0.1)(s2 + 0. 2s + 1)
and it approximation
5
g(s) =
10(s + 0.1)
where the temperature sensor and the hoist sub-systems have been replaced by two
constants.
90 Control Engineering - 1
Clearly for step inputs, u(s) = a/s, the system response is dominated by poles that lie
closest to the origin. In general, inputs excite those modes of the system that lie
closest to the poles of the input function (i.e. to the roots of Du(s)=0).
For this example, should the input to the above system be a sinusoidal function with
pole-positions close to Pole pair No 3 then the system response would be dominated
by a sinusoid.
are known as the system poles. The positions that these poles take in the s-plane
determine the performance of the system (i.e. whether it is stable or unstable,
oscillatory or damped and fast or slow)
are known as the SYSTEM ZEROS. The positions of zeros in the s-plane are also
significant, though not as characteristic as the pole positions. Zeros in fact tend to
cancel the influence of poles.
Ch 6: Prediction of System Response 91
which has one pole at s = −1 and one zero at s = −a . In the s-plane, the pole and
zero are represented by:
The time response of this system to a step input, for different zero positions is:
Note
(1) As the zero approaches the pole position from the left, it tends to
cancel the effect of the pole. In the limit, when a=1, the transfer
function becomes g(s) = 1. The zero then cancels the pole exactly and
the system has no dynamics.
(2) For step, ramp and parabolic inputs the actual response is dominated
by the pole or the zero, depending on which is nearest the origin.
(3) When the zero moves into the right-half s-plane the initial jump in the
step response is negative. System stability is determined only by the
pole position, so its response remains stable for all 'a' values.
An intuitively appealing way of viewing the effect of zeros is to expand the original
system model by partial fractions to give an equivalent parallel model:
Thus
g(s) = g1 (s) + g 2 (s)+... +g n (s)
where each sub-system gi(s) in the parallel model consists of a single real pole or a
single pair of complex conjugate poles. The system response is
y(s) = y1 (s) + y 2 (s)+... +y n (s)
Each sub-system contributes to the overall response in direct proportion to the size
of its gain relative to the other sub-systems. And these gains are determined by the
positions of the system zeros.
Ch 6: Prediction of System Response 93
Example
Clearly the position of the zero affects the gain of the lower block and hence the size
of its contribution to the output y(t). The mathematical reasons for this are now
discussed.
n k n
= å (s − ip ==> y(t) = å k ie p i t
i=1 i) i=1
where γ is a constant, zi are the system zeros and pi are the system poles.
m n
= γ ∏ (p j − z i ) / ∏ (p j − pi )
i=1 i≠ j
In the s-plane these vectors for a simple system (in which complex conjugate pairs
have been omitted for clarity) are:
(1) The closer the zero moves to pole j the shorter vector a becomes and
hence the residual kj associated with pole j decreases.
(2) The closer the other poles move to pole j the shorter vector b becomes
and hence the associated residual kj increases.
7
Routh-Hurwitz Stability
Criterion
Clearly it is important that the poles of a system lie in the left-half s-plane since this
ensures that it has a stable response. One method for determining how many
unstable poles a given system has is the ROUTH-HURWITZ CRITERION.
where N(s) and D(s) are co-prime polynomials in s. (Co-Prime means that all factors
common to N(s) and D(s) have been cancelled. It is an important constraint on the
polynomials of g(s), even though it is not often stated.) Once again g(s) can
represent either an open loop or a closed loop configuration as required.
Stability of the system is determined by its pole positions in the s-plane (i.e. by the
location of the roots of its characteristic equation
From the theory of polynomials, inspection of the coefficients, bi, of D(s) can
indicate stability of transfer function g(s) since:
Rule No 2: If any coefficient bi, except bo, is zero then g(s) has
unstable poles or oscillatory poles.
Further information about the poles of g(s) can be obtained by constructing and
analysing the Routh array:
96 Control Engineering - 1
where the first two rows contain coefficients of the original polynomial D(s) and the
subsequent rows (3 to n+1) are computed by the determinants
1 bn b n − k −1
cn − k = −
b n −1 b n −1 b n − k − 2
1 b n −1 b n − k − 2
dn−k = − etcetera.
c n −1 c n −1 c n − k − 2
Two special cases can arise in computing the Routh array, namely:
(1) if a zero appears in the first column and there are non-zero elements in
that row then replace the zero with an algebraic variable, ε (>0), and
continue;
(2) if a complete row of zeros appears then this row must be replaced by
the derivative of the previous row (as shown in the examples).
The completed table is then analysed according to the Routh criterion which states
that:
Rule No 3: The system, g(s), is stable iff all elements in the first
column are strictly positive (>0).
Example 1
Initial conclusions
Ch 7: Routh-Hurwitz Stability Criterion 97
s4 1 6 2 7
s3 2 3 1 (0) ← Pad row with a zero
s2 3 ε 7 ← Replace 0 with ε
s 4 (ε-21)/ε (0) ← Pad row with a zero
1 5 7
Conclusion. There are two sign changes so the system has two unstable poles. (This
is confirmed since D(s) has roots at s = −0.82 ± j0. 75 and s = 0 . 57 ± j0.80 )
Example 2
If all the elements in a row are zero, then the polynomial d(s) has an even factor with
coefficients given in the row above. The table may be continued by replacing the
row of zeros with the coefficients of the derivative of the even factor.
s5 1 1 3 2
s4 2 4 12 8
s3 3 0 0 ← Entire row of zeros
The row full of zeros implies an even factor in the previous row
a(s) = 4s 4 + 12s2 + 8
Replace the row of zeros with coefficients of the derivative of the even factor:
98 Control Engineering - 1
d
a(s) = 16s3 + 24s
ds
s5 1 1 3 2
s4 2 4 12 8
s3 3 16 24 (0) ← Pad with zero
s2 4 6 8
s 5 8/3 (0) ← Pad with zero
1 6 8
The first column is [1 4 16 6 8/3 8]T. No sign changes indicate that there are no
unstable poles in this characteristic function. (Its poles are at s = −4 , s = ± j1. 4 and
s = ± j1. 0 )
Assume that the controller constants, K and T, are set by potentiometers that plant
personnel can tune. Use the Routh array to find ranges of K and T that ensure a
stable system response.
The transfer function model, q(s), for the open loop is:
Ch 7: Routh-Hurwitz Stability Criterion 99
s3 1 T KT
s2 2 T K
s 3 K(T-1) (0) ← Pad with zero
1 4 K
This illustrates how the Routh Table can be used to determine the range of
parameter values over which a system will remain stable. Such information is
conveniently presented in a DESIGN DIAGRAM that summarises the control
engineering conclusions in a general format that is readily understood by others.
Use the modified Routh Array to determine whether the system will settle to within
±2% of its final value in less than 20 [s].
which shifts the Y-axis 0.2 units (from 4/20) to the left in the s-plane. Clearly
instability in the v-plane translates directly to a system that is too slow (or unstable)
in the s-plane. Graphically:
Since one coefficient of this polynomial, dv(v), is negative there is at least one pole
that is too slow (i.e. to the right of the s = −0. 2 line)
Ch 7: Routh-Hurwitz Stability Criterion
101
v3 1 1 17.2
v2 2 9.5 -1.80
v 3 17.4 (0)
1 4 -1.80
There is one sign change in the first column so only a single root of the original
characteristic equation is too slow. (The roots of D(s), found by numeric methods,
are at s = −7. 2 , s = −2 . 8 and s = −0 .10 )
Once again the Routh-Hurwitz extension is probably most useful when the
coefficients of the original D(s) are unknowns. Unfortunately the resulting algebraic
manipulations tend to be extremely laborious.
The simple translation of the jω axis in the s-plane to form the v-plane can be
generalised using Mobius transformations. As an example, a circular region in the s-
plane that includes damping and transient decay rate is mapped to the left-half of a
v-plane by the transformation pair
γ +v β s + αβ − 1
s= −α and v=γ
β ( γ − v) β s + αβ + 1
So far transfer function models have been dealt with in some depth to illustrate:
Experience dictates how many of these criteria are met. Common closed loop
configurations and design aims are described below to provide some guidance.
Combining these equations to eliminate the internal variables, e(s) and u(s), gives
y(s) = g(s) k(s) e(s)
= g(s) k(s) [r(s) - y(s)]
104 Control Engineering - 1
and
h(s) = [1 + q(s)]−1 q(s) Closed loop model: y(s) = h(s) r(s)
Thus setpoints are tracked accurately whenever the open loop model, q(s), is large,
irrespective of the actual plant model, g(s). Pause and take careful note of the
design requirements that are contained in this closed loop analysis:
This is a clear goal that is elegant, practical and easy to strive for, provided it is
remembered.
Large open loop gains can usually be realised exactly at low frequencies but fall off
at higher frequencies. Obviously instability will limit the size that k(s) can attain in
practice, as will signal amplitudes, actuator saturation, noise levels, etc. (Large gains
also ensure that the closed loop is insensitive to disturbances and changes in process
behaviour.)
The role of the control engineer is to balance these factors to produce an acceptable
design for the given situation.
Ch 8: Feedback Control Systems 105
Example
A simple numeric case illustrates the practical importance of high gain in feedback.
Consider a non-linear, time-varying plant that has been modelled simply by a gain
g(s) = 5 (±4)
where the nominal model, 5, is roughly equal to its uncertainty, 4, and represents a
±80% variation in gain. A unity feedback configuration for the process, which
includes a controller, k(s), that is also merely a gain, ensures that variations in the
closed loop gain are reduced significantly as indicated:
Clearly the closed loop configuration itself ensures considerable improvement over
the open loop process, and the gain of h(s) becomes more consistent as loop gain is
increased. (This phenomena holds for more complex transfer function models, but is
less obvious. The manner in which loop gain is limited by instability and constraints
on the process input, does however require a more elaborate transfer function
approach, and is not pursued here.)
Here the model for the closed loop system is derived from the equations
106 Control Engineering - 1
And so, eliminating the internal variables u(s) and e(s), the equations become
q(s)
y(s) = r(s)
1 + q(s) f (s)
= h(s) r(s)
1
→ r(s) when qf(s) >> 1
f (s)
Note that feedback can be used to invert functions (e.g. Convert capacitors to
equivalent inductors, differentiators to integrators, multiplication to division, etc.).
Example
F
K 0.1 1 10
0.1 0.20 0.17 0.07
1 1.67 0.67 0.10
10 6.67 0.95 0.10
Table of h(s)
Observe that the closed loop model approaches 1/F as gkf(s)=qf(s) increases above
one. In practical applications, it is vital to determine where exactly in the loop the
gain is situated since this has far reaching implications for the system components.
The theory presented here merely indicates the necessary goals.
Ch 8: Feedback Control Systems 107
is a ratio of two polynomials, N(s) and D(s), then the closed loop system h(s) has the
Characteristic equation
D(s) + N(s) = 0 ... (8.1)
N q (s)D f (s)
since h(s) = [1 + q(s)f (s)]−1 q(s) = .
D(s) + N(s)
The roots of Equation (8.1) define the closed loop pole positions and hence the
performance of the feedback system.
Note
all have roots in the s-plane in the same positions as the Characteristic
equation for the closed loop system. (This fact is exploited later.)
(2) For unity feedback systems, the closed loop zeros of h(s) are the same
as the open loop zeros of q(s). Thus unity feedback only alters pole
positions.
(3) In addition to modules k(s) and f(s), the designer may want to include
a pre-filter module, p(s), on the closed loop setpoint. This has no
108 Control Engineering - 1
The design of k(s) involves the informed manipulation of differential equations and,
without suitable theoretical tools, presents a difficult, if not impossible problem to
be solved. Many design techniques have been developed over the years to deal with
various classes of design problems and the control engineer needs to know which of
these methods is most appropriate in any given situation, what its advantages and
disadvantages are, and how the methods can be mixed to deal successfully with
difficult practical problems.
In the following chapters two distinctly different approaches to the design of control
schemes are considered:
The dynamic part, q*(s), of the open loop transfer function is written as
110 Control Engineering - 1
N* (s)
q * (s) =
D* (s)
where the coefficients of the highest power of s in both numerator and denominator
polynomials are forced to unity. The ROOT LOCUS GAIN, γ, is a constant gain
factor that is taken to be positive. (Root loci can also be drawn for negative gains γ,
but these are not considered here.)
Note
(1) The open loop transfer function q(s)=g(s)k(s) models the cascade
connection of the plant model, g(s), and the controller model, k(s).
(2) The open loop transfer function, q*(s), used in the Root Locus is a
scaled version of the actual open loop transfer function, q(s).
(3) The Root Locus Gain is not actually a gain in the traditional sense,
though it is proportional to the open loop gain. It is very useful for
reading gain values directly from Root Locus diagrams.
γN* (s)
h(s) =
γN (s) + D* (s)
*
The roots of this equation give the positions in the s-plane of the closed loop poles.
Also the closed loop zeros of h(s) are exactly the same as the open loop zeros of
q(s). i.e. Unity closure of the feedback loop in single-variable systems only alters the
pole positions of the overall system.
Now, for any given value of the constant γ, the closed loop poles of h(s) can be
computed from its Characteristic Equation (Equation 9.1) and plotted in the s-plane.
As the root locus gain, γ, is changed from 0+ to +∞, these closed loop poles move
around in the s-plane, from the positions of the open loop poles to the positions of
the open loop zeros, thereby tracing out the ROOT LOCI. (Thus the root locus
Ch 9: Root Locus Design Method 111
An Illustrative Example
which has a gain of 2 and two time constants 0.2 and 0.1 [seconds]. The open loop
transfer function q(s) has two poles, one at s = −5 and one at s = −10 . It also has no
zeros (which is equivalent mathematically to having zeros at infinity).
The scaled transfer function, q*(s) required for drawing the Root Locus is derived as
follows
100 1
q(s) = = 100 = 100 q* (s) = γ q * (s)
(5 + s)(10 + s) (s + 5)(s + 10)
In a unity feedback configuration this open loop model generates the closed loop
characteristic equation
s2 + 15s + (50 + γ ) = 0
For this simple example the closed loop poles can be computed analytically and are
15 1
s=− ± 152 − 4(50 + γ )
2 2
Clearly the closed loop pole positions are a non-linear function of the root locus
gain γ
s = s( γ )
As γ is varied, this function sketches the root loci in the s-plane for this process
model
112 Control Engineering - 1
Notice that the closed loop poles start at the positions of the open loop poles (i.e. at
s= −5 and s= −10 ) when γ=0. Also the closed loop poles move off to the zeros at
infinity as γ tends to infinity.
Knowing the significance of pole positions in the s-plane, the Root Locus, once
drawn, is studied to predict the dynamic performance of the closed loop system.
(2) Its response is dominated by a slow C/L pole when γ << 6.25.
(5) The fastest transient decay rate is achieved at γ=6.25 (The ±5% band
is entered in 0.40 [sec] and the ±2% band in 0.53 [sec]).
(6) The transient decay rate does not change for gains γ > 6.25.
These predictions are confirmed by its response to a step change in the setpoint for
different gains:
Ch 9: Root Locus Design Method 113
Note
As expected the error between the setpoint and output reduces when γ is increased.
Unfortunately its closed loop performance deteriorates since the controlled system
also becomes more oscillatory. The input also increases in amplitude.
(1) Symmetry
The root locus diagram is symmetric about the real axis. (Obviously)
114 Control Engineering - 1
(2) Origin
Each locus originates at an open loop pole when γ=0. Thus there are <n>
loci where 'n' is the order of the open loop system q(s).
(3) Terminus
When γ → ∞, <m> of the <n> loci terminate at the <m> zeros of q(s),
while <n-m> loci approach infinity along asymptotes.
(4) Asymptotes
(180 ± 360k)
where k is an integer.
(n − m)
Complex poles and zeros of q(s) have no effect on the loci that follow the
real axis. Segments of the real axis are part of the root locus iff the total
number of real poles plus real zeros to their right is odd.
The locus leaves an open loop pole at an angle determined by a trial point,
close to the pole, satisfying the angle condition:
This rule is deduced from the fact that the root locus gain, γ, is a real
positive constant given by
Ch 9: Root Locus Design Method 115
D* (s)
γ =−
N* (s)
and that the angle, ϕ(γ), of the root locus gain, γ, must be zero. This
condition yields the equation
These are also found either (a) by using the angle condition (Rule 6) for a
trial point chosen close to the real axis, or (b) from the differential
d
γ (s) = 0
ds
so that
d d
N(s) D(s) − D(s) N(s) = 0
ds ds
Example
To illustrate how the rules for sketching Root Locus are used in practice, find the
value for the controller gain K that give pole positions which ensure a damping
factor of 0.7071 for the closed loop system:
2K
q(s) =
s(1 + 0. 2s)(1 + 0.1s)
The scaled transfer function needed for the root locus is given by the relationship
100K 1
q(s) = = 100K = γ q * (s)
s(s + 5)(s + 10) s(s + 5)(s + 10)
Here
<n>=3 as there are 3 open loop poles ( s = 0 , s = −5 and s = −10 ).
<m>=0 as there are no zeros (or all zeros at infinity).
so
<n-m>=3.
The open loop poles, marked as X, are plotted on the s-plane. These give the starting
points on the root loci (where γ = 0). Any zeros, marked as O, are also plotted
and indicate the termination points (Where γ→+∞). In this example there are only
poles and the loci move along asymptotes to terminate at infinity.
The Rules that are activated to plot the root locus are:
Rule 2
There are <n> or 3 loci, starting at s = 0 , s = −5 and s = −10
Rule 3
The <n-m> or 3 loci approach infinity along 3 asymptotes
Rule 4
The asymptotes are angled at (180±360k)/3 = 60° ± 120k = ±60°, 180°
The intersection point σ c = {(0) + ( −5) + ( −10)} / 3 = −5
(These can now be drawn in the s-plane)
Rule 5
The loci on the real axis are simply drawn as shown:
Ch 9: Root Locus Design Method 117
Rule 8
The locus break-away point on the real axis is computed from a trial point
near the axis, as shown in the following sketch.
Rule 6
The angle criterion for the trial point is
ϕ1 + ϕ 2 + ϕ 3 = π ± 2kπ
From the geometry of the trial point, the angle conditions translates to
δ δ δ
[ π − tan−1 ( )] + [ tan−1 ( )] + [ tan−1 ( )] = π ± 2 kπ
−σb σ b − [ −5] σ b − [ −10 ]
δ δ δ
π+( )+( )+( ) = π ± 2 kπ
σb σb + 5 σ b + 10
And so
δ δ δ
( )+( )+( )=0
σb σb + 5 σ b + 10
3 σ 2b + 30 σ b + 50 = 0
or
σ b = −2.11 (σ b = −7. 89 is discarded. It holds for γ ≤ 0)
The remaining complex portion of the Root Loci are now sketched (requiring a little
artistic talent) to give the complete root locus diagram for the system.
The closed loop pole positions (marked as small boxes) satisfy the specification of a
damping factor of 0.7071 for the controlled system h(s). The Root Locus Gain, γ, at
this point, s0, is computed from the magnitude relationship
D* (s0 )
γ=
N* (s0 )
n
∏ (s0 − pi )
= i=1
m
∏ (s0 − zi )
i=1
And so the controller gain that gives a closed loop system with a damping factor of
0.7071 is:
Ch 9: Root Locus Design Method 119
83. 7
K= = 0.837
100
Note
From the Root Locus diagram for a closed loop system, it is possible to find values
of γ that ensure particular closed loop pole positions on the given loci.
The engineering problem is how to position poles and zeros of a controller, k(s), in
the s-plane to improve on the root loci (i.e. on the closed loop response).
To design a compensated closed loop system, start by plotting the root loci for the
plant transfer function (that includes a component to ensure the correct type
number). These loci show closed loop pole positions for the uncompensated system
and also present the problem in an appropriate form that gives engineering insight
into (a) the dynamic performance of the closed loop system and (b) how to alter the
open loop system to improve on its closed loop performance.
The following rules of thumb are useful in getting started (though not always
correct):
(3) closed Loop poles close to the origin dominate the response (unless
cancelled by a nearby zero),
(4) closed Loop zeros are identical to the Open Loop zeros (for unity
feedback),
Example 1
(2) Transients that decay to the ±2% band within 2.5 [seconds].
The region in the s-plane within which the closed loop poles should lie is given by:
Example 2
The root locus for the open loop system, g(s), shows where the poles of the closed
loop system move as the open loop gain is increased.
The root locus diagram shows quite clearly that closing the loop on the plant g(s),
with unity feedback and no compensation, results in a closed loop system that is
unstable for all values of Root Locus Gain, γ. (Because there is always one pole and
sometimes two poles of the closed loop transfer function
h(s) = g*(s) / [1+γg*(s)]
Putting a stable open loop zero near the origin will attract the loci towards (and
hopefully into) the stable left-half s-plane. For example, a cascade element
s + 0. 5
k(s) =
s+3
This compensator is a LEAD CIRCUIT. It does not violate the implicit properties
of a satisfactory control element since it is
(3) Has a dominant zero at s = −0 . 5 which should attract the loci towards
the left-half s-plane.
As the loop gain is increased, the three closed loop poles move along the root loci.
The unstable closed loop pole on the loci starting at s=+2 is attracted towards the
zero of the compensator, k(s), and crosses into the stable left-half s-plane when the
root locus gain, γ=12. This value for γ is computed from the product formula:
Product of distances to all poles
γ =
Product of distances to all zeros
(3)(1)(2)
= = 12
(0. 5)
Obviously the highest possible loop gain will merely result in a transient decay rate
determined by the dominant and complex conjugate poles at s = −0. 75 ± jb so there
are going to be limits on the transient decay rate that can be attained with this
control system. (This can be improved by further additions to the compensator.)
As expected, the zeros of h(s) are the same as those of q(s). The closed loop pole
positions are given by the Characteristic Function of h(s)
or, explicitly,
s=0 and s = −1 ± j 6
From the root locus it is obvious that a slightly higher loop gain is needed to
produce a stable closed loop system. The step response of the original unstable open
loop plant and a stabilised closed loop for γ=20 is shown below, together with the
input required to achieve stabilisation.
124 Control Engineering - 1
Observation
Clearly the root locus diagram for g(s) has guided the designer towards a dynamic
compensator k(s) that stabilizes the system in a closed loop configuration. Further
refinement in the constants of the present k(s) or perhaps a few more zeros (and
poles) should be tested before coming to a final decision.
(Clearly the Type number could do with some improvement, but will destabilise the
loop considerably.)
Consider the problem of designing a range of I values that can be used to trim the
following closed loop system without causing instability.
Ch 9: Root Locus Design Method 125
and so
4(1 + sI)
h(s) =
4(1 + sI) + sI(1 + 10s)
The Characteristic equation for the closed loop system, h(s), can be written as
0 = 4(1 + sI) + sI(1 + 10s) = 4 + Is(5 + 10s) = 4 + 10Is(s + 0. 5)
1
= + (s2 + 0. 5s) (Dividing by 10I)
2. 5I
= γ + s(s + 0. 5)
This equation has the polynomial form required for drawing Root Loci. It has Open
Loop Poles at s = 0 and s = −0 . 5 and no zeros.
(1) The closed loop system is stable for all values of γ, and hence I since
γ=1/2.5I.
(2) At a damping factor of 0.7071, the Root Locus Gain is measured from
the diagram and found to be
γ=(0.354)2=0.125
1 1
It has one zero at s = − and one pole at s = − . In the s-plane it provides the
T αT
pole-zero pair:
If the input and output of the compensator are in volts then the lag function is
implemented by the passive electronic circuit:
1 + sR 2 C
= [V]/[V]
1 + s(R1 + R 2 )C
In Root Locus designs the dominant pole is used to repel loci in such a way that the
response is improved while the zero helps retain closed loop stability.
1 1
It has one zero at s = − and one pole at s = − . It provides the pole-zero pair:
αT T
R2 1 + sR1C
= [V]/[V]
R1 + R 2 1 + s R 2 R1 C
R1 + R 2
In Root Locus designs, the dominant zero is used to attract loci in such a way that
the closed loop is stabilized. The pole prevents excessive noise transmission.
Note
The Lead circuit tends to amplify high frequencies, so its response is generally quite
noisy. The problem is aggravated by large α-values, hence the reason for its upper
limit of 10. In practice it may be better to use two lead circuits in cascade, with
small α-values, than one lead with a larger α-value.
For large industrial systems a process control computer is often available so the
compensator k(s) can take the form of any stable transfer function
a 0 + a1s + a 2s+... +a ms m
k(s) = with m ≤ n
b 0 + b1s + b 2s+... +b nsn
Analog electronic circuits that realise such transfer functions can also be designed as
discussed in Chapter 17.
Whatever the requirements, the final choice should be biased towards the simplest
possible k(s) that will do the job (as the ultimate goal in engineering is simplicity).
10
Dead Time and the Root
Locus
The ROOT LOCUS diagram is a very powerful analytical method for dealing with
dynamic systems. Unfortunately it is not universally applicable and is deficient in
some important features. For example:
(1) It is not able to deal easily with all types of dynamic systems that are
found in practice, particularly in industrial application of control
engineering.
(4) It has not yet been extended to deal effectively with complex
multivariable processes commonly encountered in industrial
applications of advanced microcomputer control schemes.
The first point is now dealt with in detail to motivate a different approach to system
modelling and controller design.
The time lapse between the instant when the water valve is stepped and the time
when the density starts to respond is known as DEAD-TIME, τ (Tau). (Its value is
clearly dependent on flowrate but is assumed to be constant during analysis.)
From the Real Translation law for Laplace transforms, the response of the system
can be modelled by
y(s)
= g(s) e − sτ [kg / m 3 ] / [o ]
u(s)
a 0 + a1s+... +a ms m
g(s) = with m ≤ n
b 0 + b1s+... +b nsn
describes the dynamic response of the process, while the exponential term, e − sτ ,
accounts for its dead-time.
132 Control Engineering - 1
The dead-time term is extremely difficult to handle by Pole-Zero methods, like Root
Locus, since it effectively introduces an infinite number of poles and zeros into the
system transfer function.
e −sτ /2 e−θ sτ
e − sτ = +sτ /2
= +θ
defining θ =
e e 2
1 1 1 1
1 − θ + θ2 − θ3 + θ 4 − ... +(-1)n θ n +...
= 2! 3! 4! n!
1 1 1 1
1 + θ + θ2 + θ3 + θ 4 +... + θ n +...
2! 3! 4! n!
Note
(2) Historically the Pade approximations were, and probably still are, very
important for analog electronic implementations of dead-time terms.
The frequency response models contain neither poles nor zeros explicitly, yet their
mathematical descriptions can predict implicitly pole/zero positions. In this way
frequency response methods are used for the design of feedback controllers.
where u(t) = uosin(ωt) is the input and y(t) = yosin(ωt+φ) is the output response
(after all transients have decayed).
134 Control Engineering - 1
y0
The AMPLITUDE RATIO (ω )
u0
Alternatively, transfer function models are readily derived from an analysis of step
test data that is easily obtained, and frequency response models can be derived from
transfer function models by a change of variable. The relevant theory follows.
ω u0
= (From Laplace Tables)
s + ω2
2
kc k *c
y(s) = + + Transients due to g(s)
(s − jω ) (s + jω )
Ch 11: Frequency Domain Design Methods 135
where the residuals k c and k *c are complex conjugate constants, evaluated from the
limits
ωu 0
k c ( jω ) = lim [(s − jω ) g(s) ]
s→ jω s2 + ω 2
ωu 0 u
= lim [ g(s) ] = g( jω ) [ − j 0 ]
s→ jω (s + jω ) 2
and
ωu 0 u u
k *c ( jω ) = lim [ g(s) ] = g( − jω ) [ j 0 ] = g* ( jω ) [ j 0 ]
s→− jω (s − jω ) 2 2
g( jω ) = ρe jθ
which can be computed directly from the transfer function model, g(s).
Thus the Frequency response model, g(jω), is given EXACTLY by the transfer
function model, g(s), evaluated at s=jω.
136 Control Engineering - 1
Example
where y(t) is the output voltage across C in [Volts] and u(t) is the driving voltage in
[Volts]. Taking Laplace transforms, ensuring that all initial conditions are zero,
gives the transfer function model
y(s)
g(s) =
u(s)
1
=
1 + sRC
φ = Angle[g( jω )] = − tan-1 ( ω T)
12
= sin( ω t − tan−1 ( ωT) )
1 + ω 2 T2
This graph is parametrised by frequency, ω. Each point on the plot is given in polar
co-ordinates by the vector (<y0/u0> , <φ>).
138 Control Engineering - 1
Note that the frequency response model is easily derived from either the original
differential equations or the Laplace transforms by direct substitution of (jω) for
d
( ), (D) or (s). As before, it is assumed that all initial conditions are zero.
dt
Example 1
(1 + 4 D + D2 ) y(t) = (2 + 3 D ) u(t)
where (jω) has been substituted for the derivative operator (D).
Example 2
e −3s
y(s) = u(s)
1 + 7s
e − j3ω
g( jω ) =
1 + j7 ω
1
= g1 ( jω ) e − j3ω where g1 ( jω ) =
1 + j7 ω
Note
− j3ω
(1) The dead-time term, e , is merely the complex number
(2) The dead-time term, e − j3ω , does not change the magnitude of the
frequency response model, g1(jω). It only increases its phase angle by
−3ω [radians].
Recall that the stability of a closed loop control system is determined by its pole
positions in the s-plane. Although the frequency response model no longer contains
explicit information on system poles, particularly when dead-times are involved, it is
possible to study a polar plot of the model of an open loop system, q(jω), and to
predict whether the poles of its corresponding closed loop system, h(s), are stable or
not. (The analysis is based on the Principle of the Argument.)
The polar plot of a frequency response model, q(jω), is simply its complex value
plotted on an Argand diagram as a function of frequency, ω. (The Argand diagram
for q(jω) is called the q-plane in these notes.)
2
q(s) =
1 + 5s
Graphically the complex numbers 's' and 'q(s)' can be shown as two 2-Dimensional
vectors on separate Argand diagrams, or complex planes
Clearly, as indicated on the figure, function q(s) can be considered to have mapped
the point 's = 0+j2' in the s-plane to the point 'q = 0. 2 e − j84.3 ' in the q-plane.
0
Extending this concept further, the function q(s) can be evaluated and plotted in the
q-plane as the variable 's' traverses its imaginary jω-axis in the s-plane, from ω=0
to ω→∞ . This operation can be viewed as mapping the entire jω-axis in the s-plane
to a contour in the q-plane. Obviously the points that q(s) maps out as 's' traverses its
imaginary axis are actually the points q(jω) and correspond exactly to the points
drawn out by its frequency response model.
Thus the complex function q(jω) can be viewed as a mapping of the jω-axis in the s-
plane onto a contour in the q-plane. A typical result is the polar plot:
Ch 11: Frequency Domain Design Methods 141
It is important to notice that the vector from the point (-1/f , 0) to the point q(jω) in
the polar plot represents the complex number
1
+ q( jω ) where f is a real-valued constant.
f
And, if a new set of axes were constructed to pass through the (-1/f , 0) point, then
the given polar plot also represents the frequency response of the complex function
1
X(s) = + q(s)
f
Now, let 's' take on more values by moving clockwise along a large semi-circle of
infinite radius until it reaches the s = − jω axis. Finally let 's' move up this negative
axis until it returns to its origin s = 0. In this way 's' traverses a huge semi-circle that
encloses the whole of the right half of the s-plane. The path that 's' followed is
important for control and is referred to as the NYQUIST CONTOUR or the D
CONTOUR.
Contour D is designed to enclose all the right-half of the s-plane and to avoid any
singularities on the jω axis due to poles of X(s) by indenting into the right-half s-
plane as shown, when necessary. (These poles are same as those of q(s).)
A plot of X(s) itself, as s traverses the Nyquist contour, traces out a contour in its
own plane. Note that the shape of the contour of X(s) is identical to that of q(s).
Also the origin of the X-plane is the so-called CRITICAL POINT at (-1/f, 0) in the
q-plane. A typical X contour and its q-plane equivalent are:
If X(s) is a complex function of 's' with both zeros and poles in the s-plane, and
<Z> = The number of zeros of X(s) that lie inside the D contour
<P> = The number of poles of X(s) that lie inside the D contour
<N> = The net number of encirclements that X makes of its origin
Example
In the above plots, it is obvious that <N>=2 since X(s) encircles its origin twice.
Exactly the same is concluded by studying q(s), which encircles its critical point at
(-1/f , 0) twice. This means that
2= <Z>−<P>
and X(s) has two more zeros than poles in the right-half s-plane. (The precise
Ch 11: Frequency Domain Design Methods 143
number of zeros and number of poles is not given by the principle of the argument.)
Note that the contour X(s) and q(s), as 's' traverses the full Nyquist contour, is
symmetric about its real axis. Thus it is common practice only to evaluate and plot
q(s) as s traverses from point 'a' to point 'c' through point 'b' in the D contour
Also when 's' traverses from point 'b' to point 'c' the open loop transfer function
a ms m
→ →0 if m < n
b ns n
am
→ if m = n
bn
In both cases, moving 's' from point 'b' to point 'c' has no effect on q(s) near its
critical point, and can be ignored (unless a n = − b n ).
Hence the polar plot used in practice shows only q(jω) as 's' moves from point 'a' to
point 'b' on the Nyquist contour. This is simply the Frequency Response model
144 Control Engineering - 1
plotted on polar co-ordinates and is known as the NYQUIST PLOT of q(s). The
contour X(s) is immediately deduced from q(s).
N(s)
q(s) = g(s) k(s) =
D(s)
where φc (s) is the closed loop characteristic function and φo (s) the open loop
characteristic function of the feedback system:
Element f is a constant gain and the origin of X(s) is at (-1/f , 0) in the q-plane.
The principle of the argument, applied to this X(s), holds that the net number of
likewise encirclements that q(s) makes of its critical point (-1/f , 0) is given by
By inspection of the Nyquist plot of any open loop system model, q(s), it is possible
146 Control Engineering - 1
to count the number of likewise encirclements, <N>, that q(jω) makes of its critical
point (-1/f , 0) as 's' traverses contour D. Then its closed loop system, h(s), is stable
if and only if
< N > = − p0
where p 0 is the <No of unstable O/L poles>. This criterion ensures that the number
of unstable closed loop poles, <Z>, is zero (i.e. h(s) is stable).
Example 1
Find all values of the feedback gain, f, that ensure stability of the closed loop
system:
From q(s) it is clear that p 0 = 0 because there are no unstable poles in the open
loop. (Determined from the Routh-Hurwitz Array, if necessary.) Thus for C/L
stability the number of encirclements of the critical (-1/f, 0) point, <N>, must be
zero.
which means that the closed loop system is going to be stable for all gains in a
negative feedback configuration. However it is interesting to note that the closed
loop is also going to be stable for a limited range of positive feedback gains, namely,
for
1 1
− <f<0 (from − > 4 with f < 0)
4 f
(In this case positive feedback is not recommended, even though it is stable, since
the resulting closed loop system will be slower than the open loop.)
Example 2
Find values of controller gain, k, that stabilise the open loop system q(s) in the
feedback control loop:
In analysing this loop it is generally useful to move the controller gain, k, into the
feedback path by block diagram algebra and to study the modified system which has
a closed loop response given by
k q(s)
y(s) = r(s)
1 + k q(s)
148 Control Engineering - 1
q(s)
= f r(s) where f = k
1 + f q(s)
= h(s) r(s)
Movement of the controller gain into the feedback path allows the theory to
incorporate a mobile critical point ( −1 / f , 0) into the Nyquist plot. This critical
point is readily moved along the real q(jω) axis until a satisfactory gain is achieved.
Once the design is complete, the controller gain is then merely moved back into the
forward path for controller implementation.
this line write (+1) if the crossing is in a likewise direction to that being traversed by
's' along the D contour or (-1) if it is in an anti-likewise direction. To ascertain the
value for <N>, total the (+1) and (-1) of the diagram.
For example, consider the complicated (but NOT atypical) Nyquist plot:
is generally of more interest that absolute stability. This means that designers like to
know whether or not the poles of the closed loop system lie within the following
region shown shaded in the s-plane where Zeta, ζ, is the damping factor for second
order systems.
(1) PHASE MARGIN, that lets Nyquist plots of the open loop model, q(j
ω), indicate whether the poles of the closed loop system, h(s), are
damped or not,
These two parameters measure the amount of extra phase angle and extra gain for
q(jω) that would result in marginal stability of the closed loop system (i.e. that
would give closed loop poles along the imaginary axis in the s-plane).
A few examples that illustrate how the Phase Margin and Gain Margin are measured
from particular Nyquist Plots are now considered.
By definition of the phase margin, it is measured directly from the Nyquist diagram
and found to be
PM = +24o
From the Nyquist Plot the Gain Margin (GM) is defined by:
1
GM = − with f > 0 always, and a < 0 here.
af
1
= 20Log10 [ − ] in [dB]
af
Ch 11: Frequency Domain Design Methods 153
Observe that multiplying the open loop transfer function by the gain margin forces
the Nyquist plot of q(jω) to pass directly through the critical point and results in a
closed loop system that oscillates forever. Any further increase in gain would make
the closed loop system unstable.
The plot does not cross the negative real axis in the q-plane and so
GM = ∞
11.6.1. Closed Loop Damping Factor from Open Loop Phase Margin
A number of useful relationships between the phase margin of an open loop system,
q(s), and the performance of its closed loop system, h(s), are now derived.
Assume that the closed loop system is approximated by a dominant second order
response with unity gain. Its transfer function is
ω 2n q(s)
h(s) = =
s 2
+ 2 ζω ns + ω 2n 1 + q(s)
ω 2n
=
s(s + 2 ζω n )
where ζ is the damping factor of h(s) and the Phase Margin of q(s) is given in
degrees [°]. The approximation holds good for the range
0 < ζ < 0.7071
0 < PM < 64°
Example
For a closed loop system to have a damping factor of 0.6 or more, the phase margin
of q(jω) must exceed 60°. This provides a design constraint on the Nyquist plot.
the maximum percentage overshoot of y(t) above its steady state value is given by:
Ch 11: Frequency Domain Design Methods 155
− πζ
PercentageOvershoot, M p , of h(s) = 100 exp ( )
1 − ζ2
Example
A closed loop system, h(s), in which the open loop system, q(s), has a phase margin
of 27° will have an overshoot of approximately 48% when its setpoint is stepped.
For higher order systems only the general trends hold true, unless a dominant pole-
pair exists. Many industrial engineering applications exhibit first or second order
behaviour, so these approximate relationships are extremely practical.
In the Nyquist diagram this identifies the graphical criteria that frequency response
models of open loop plants should lie to the right of the points indicated in the
following diagram:
A Nyquist diagram for the open loop frequency response model q(jω) can be used to
predict exactly the gain |h(jω)| of the frequency response model of its closed loop
system, where:
and
q(s)
h(s) =
1 + q(s)
x + jy
= if q(jω) = x + jy
1 + x + jy
M2 M2
(x + )2 + y 2 =
M2 − 1 (M 2 − 1)2
M2
with Centre (− , 0)
M2 − 1
M
and Radius
M2 − 1
Ch 11: Frequency Domain Design Methods 157
On the Nyquist diagram a number of constant M-circles can be drawn for different
values of closed loop gain M.
Thus it is possible to determine the amplitude ratio of the frequency response model
h(jω) of a closed loop system by inspection of the Nyquist plot of its open loop
frequency response model q(jω), on which constant M-circles have been drawn.
Example
To predict the bandwidth of the CLOSED LOOP system, h(s), from a Nyquist
diagram of its OPEN LOOP model, q(jω) observe the frequency at which the plot of
q(jω) crosses into the M=0.7071 circle which marks the -3dB point of the closed
loop frequency response. (It is assumed that h(j0)=1, or equivalently that q(j0) lies
on an M≈1 circle.)
When the Nyquist plot is used to design a compensator k(s) for a process model g(s)
the maximum M-value that q(jω) reaches should be limited to the range
1.1 < Mmax < 1.5 or 1 dB < Mmax < 3 dB
to avoid excessive oscillations in the closed loop system response. This provides an
alternative specification for designs using Nyquist Plots.
158 Control Engineering - 1
1 1 N2 + 1
(x + )2 + (y − N)2 =
2 2 4N 2
where N = tan(φ).
A design session starts by plotting the frequency response model, g(jω), for the plant
(plus any type number adjustment) on a Nyquist diagram. This assumes that k(s)=1
and presents the problem, g(s), in an appropriate form.
To continue, the following design information needs to be given in the same form
(i.e. on a Nyquist plot):
(1) the closed loop specifications (Phase Margin, Gain Margin, M circle,
C/L bandwidth, C/L damping factor, etc.),
The design engineer studies the Nyquist plot for g(jω) and then selects a suitable
compensator, or combination of compensators, k(s), that is likely to improve on the
original system g(jω). The new open loop transfer function q(jω) = g(jω) k(jω) is
then computed and plotted on the Nyquist diagram. After a few well-chosen
iterations a good compensator that meets the closed loop specifications is usually
found.
160 Control Engineering - 1
Example
Consider a plant:
10 e − s
g(s) = [kg/m3]/[°]
1+ s
Design a compensator for this process that results in a C/L system h(s) with a
damping factor of 0.3 or more. Also ensure an adequate Gain Margin (> 2).
First it should be noted that the dead-time in the process is of the same order of
magnitude as the time constant. Thus it cannot be neglected and the Root Locus
method is not suitable for analysing this plant.
The design calls for a damping factor of 0.3 or more in h(s). This criterion translates
to a requirement that the open loop system q(jω) have a Phase margin of 30° or
Ch 12: The Nyquist Plot Design Method 161
more, which in turn can be represented on the Nyquist plot by a line drawn through
the origin at an angle of -150°.
From the plot of g(jω) it is clear that without a compensator element, k(s), in the
feed forward path:
(1) The system g(jω) in closed loop is unstable for feedback gains, f,
greater than 0.23 (from 1/4.4).
(2) For the specified closed loop damping factor a Phase Margin of 30° is
required for the open loop model q(jω). To achieve this, the gain, f,
should be 0.19 (from 1/5.36).
(3) At a gain of f=0.19 the Gain Margin is a mere 1.22 (from 5.36/4.4) or
1.7dB and needs to be increased to exceed the recommended value of
2 or 6dB. This could be achieved by decreasing f further to 0.12 (from
1/8.8), but so small a gain gives a large damping factor, a very slow
closed loop response and worse steady state error.
(4) The steady state error at f=1 would be 0.091 (from 1/(1+10) for a
unity step setpoint), but the closed loop is unstable. At f=0.19 the
steady state error increases to 0.34 (from 1.9/2.9).
(5) The closed loop bandwidth with f=1 is in the range 3 to 4 [rad/s]
(from the frequency at which the Nyquist plot of q(jω) enters the
M=0.7071 circle). It drops to approximately 1 [rad/s] when the gain is
reduced to f=0.19.
has a low-frequency gain of 1.0 (from 1/1) and a high-frequency gain of 0.3 (from
0.6/2) for ω >> 1.4 (from 1/0.6) [rad/s].
Note that steady state error is not given much attention in this example, but cannot
be ignored in a practical situation.
The Nyquist plot of the compensated open loop system q(jω)=g(jω)k(jω) is:
162 Control Engineering - 1
The parameters '0.6' and '2' of k(s) can be trimmed, or further lag circuits could be
installed, in order to provide the necessary Gain Margin and the designer would
embark on a few more trials, each evaluated by a study of the resulting Nyquist plot
for Gain Margin and Phase Margin. (In practice Nyquist plots are computed and
displayed with ease by modern digital computers, so the design engineer's primary
task is to evaluate these plots carefully and to suggest modifications to k(s) that
improve on the characteristics of the closed loop system).
generally take the form of Lead or Lag circuits, or combinations of both. It is thus
important to know how these will affect the frequency response model, g(jω), of a
Ch 12: The Nyquist Plot Design Method 163
(2) amplify low-frequencies, which boosts loop gain and reduces steady
state errors.
It should be noted that the phase lag introduced by a lag circuit swings the Nyquist
plot clockwise towards the critical point. This is an undesirable side-effect, that is
partially minimised at high frequencies by the presence of its zero (which explains
why a simple R-C circuit having only one pole is avoided in control loops).
The maximum attenuation provided by the circuit is K/α and occurs at high
frequencies (ω > 1/T). For low-frequency amplification (at ω < 1/αT) a gain term,
K, is included in k(s). (Remember that K=1 is usually assumed since it is moved into
the feedback path to become feedback gain, f=K, giving critical point -1/f in the
164 Control Engineering - 1
Nyquist plot).
The maximum phase lag introduced by the lag circuit depends only on α and is
given by the trigonometric function
1/ α − α
φcm = tan−1 ( )<0
2
Note
(1) The phase lag is an undesirable effect introduced by the lag circuit.
In designing a lag compensator for a given plant, the initial Nyquist plot for g(jω)
would suggest what low-frequency gain is required to boost the gain of q(jω)
sufficiently to ensure that the desired closed loop gain of h(jω), indicated by M
circles, is achieved. The lag circuit gain, K, is then selected to provide this low-
frequency gain; at the same time constant α is chosen to reduce the loop gain at
high-frequencies in order to avoid instability and provide adequate gain margin. The
frequency at which the dominant effects are needed is finally set by T.
α − 1/ α
φcm = tan−1 ( )>0
2
The low-frequency gain provided by the lead circuit is K, while the undesired high
frequency gain is Kα. (Once again K=1 as this term is usually included in the
critical point (-1/f , 0) in the Nyquist plot).
In designing lead compensators, the value set for α determines the amount of phase
advance that will be introduced into q(jω) while the choice of T sets the frequency at
which this phase advance is provided. Since α also determines the high-frequency
gain, it should be kept at a minimum. (In practice, two or more lead circuits with
small α-values are more effective than one lead circuit with a large α-value.)
166 Control Engineering - 1
The Lead circuit is intended to improve on Phase Margin thus damping the closed
loop response, usually making the system faster with less oscillatory behaviour. It is
unfortunately sensitive to measurement noise.
Example
In a helicopter, lift is provided by the main rotor and manipulated by adjusting the
angle of the blades of this rotor. For the purposes of a control analysis, the dominant
relationship between <altitude> in [m] and <main-rotor pitch> in [°] of a hovering
helicopter:
This plot of g(jω) shows clearly that the uncompensated helicopter in closed loop is
marginally stable, as the damping factor of h(s) is 0. Also, since the plot of g(jω)
passes through the M=∞ circle at ω=2 the uncompensated helicopter in closed loop
will oscillate with a frequency of 2 [rad/s].
Clearly the present system, g(s), needs more phase margin. Thus a Lead circuit is
proposed to introduce phase advance into the loop in order to improve on the
damping factor of the closed loop system h(s).
Setting α=4 (initially compromising between phase advance and high frequency
gain) the Lead circuit will introduce a phase advance of
4 − 1/ 4
φ = tan−1 ( ) = 37. 0 [o ]
2
(Larger values of α would give more phase advance, but aggravate noise problems).
As an initial estimate, the phase advance should occur around ω=2 [rad/sec] (since
168 Control Engineering - 1
this is the frequency at which the magnitude of g(jω) is 1 and so the Lead circuit is
likely to have the most effect on Phase Margin). Thus because
1
ω cm = 2 =
αT
which has a Phase Margin of 32° when K=1 and an infinite Gain Margin. The
closed loop damping factor is thus 0.32 with a 43[%] overshoot to step inputs and
its bandwidth is approximately 3-to-6 [rad/s] as determined from M-circles.
Notice the large initial value (of 4 [°]) for the input. Such input spikes are common
in lead compensation controllers and should be reduced by introducing a low pass
filter into the loop. This is positioned after k(s) on signal u(s) and must have non-
dominant dynamics relative to h(s) so that it does not alter the closed loop
performance.
The Nyquist plot of q(jω) shows that a decrease in the gain, K, of the lead circuit
will increase the phase margin of q(jω) slightly and hence the damping factor of the
closed loop system, h(s). It will also slow the response of h(s) down (which is
deduced from inspection of the M=0.7071 circle).
Clearly the first attempt at designing k(s) is partially successful, but can be improved
by further adjustments that make optimum use of the phase advance provided by a
Lead circuit.
13
Alternative Frequency
Response Plots
A number of alternative representations of frequency response models have been
derived over the years. Each has its own strong features and the astute control
engineer will choose that representation which best suits the needs of the design
under consideration. Also, some industries have adopted particular representations
as their standards.
Thus the concepts derived for Nyquist Plots are now re-stated for Bode Diagrams,
Nichols Charts and Inverse Nyquist Plots.
s Pure derivative/integral
s2 + 2 ζω ns + ω 2n
A pair of Zeros/Poles
ω 2n
s
q(s) =
(1 + sT1 )(1 + sT2 )
If both {20Log10 ïq(jω)ô} and {Angle[q(jω)] in [°]} are plotted on two separate
Cartesian graphs as functions of {Log10 ω} then the magnitude graph can be
approximated quite accurately by straight line asymptotes.
Its main disadvantage is that it does not predict instability in systems that contain
any zero in the right-half s-plane i.e. in Non-Minimal Phase systems. (For such
systems the Nyquist stability criterion must be applied.) There are other
disadvantages as will become clear later.
Consider a refrigerated storage room in which the flow of Freon is used to regulate
room temperature:
172 Control Engineering - 1
y(s) 4 e −2s
g(s) = = [°C] / [litre/minute]
u(s) (1 + 5s)(1 + 60s)
Stability of the closed loop system can be determined from a Bode Diagram of the
open loop model. Specifically closed loop stability requires that:
Recall that Gain and Phase Margins of the open loop provide an indication of
robustness and damping for the closed loop system.
and its corresponding Bode Diagrams, the Gain Margin (=10) and Phase Margin
(=80°) are readily obtained from the latter:
Lag Circuit
Lead Circuit
The lead circuit improves on the phase margin for q(jω). It does however amplify
high frequency noise and should be used with caution. Its transfer function is
1 + sαT
k(s) = K 1 < α ≤ 10
1 + sT
Consider the hovering helicopter example where the transfer function model is given
by
4
g(s) =
s2
To improve on the Phase Margin, the phase plot at 2 [rad/s] must be moved upward.
176 Control Engineering - 1
A lead circuit can do this as shown by its Bode diagram. Set α=4 to give a Phase
advance of 37° and chose T=0.25 so that this phase advance peaks at frequency of ω
=2 [rad/s].
The Bode diagrams for the compensated system, q(jω) = g(jω) k(jω), are then:
Clearly a smaller value for T (i.e. a larger value for ωcm) might give better results for
the same value of α in the compensator k(s). The design for k(s) would proceed
along these lines until a satisfactory result was obtained.
It is useful to compare this example with the identical problem analysed previously
in the Nyquist plot in terms of the ease with which:
The effect that a gain factor, f, has on the design can be determined in the Bode
diagram by shifting the horizontal frequency axis in the gain plot vertically
downwards by
20Log10 f [dB]
Thus, by inspection of the Bode diagrams for q(jω) in the above example, it is clear
that reducing the gain will improve the phase margin of q(jω) and hence the
damping factor of the C/L system, though at the cost of reducing the bandwidth.
Bode Diagrams CANNOT be used to test for stability of such systems as the results
are invalid. For these systems the Nyquist stability criterion MUST be used.
Example
which has an unstable pole at s=+1 and is thus non-minimal phase. In closed loop
this system is unstable for all gains, f ≥ 0, as can be seen using the Nyquist stability
criterion.
the phase angle never reaches -180° and so the closed loop system appears to be
stable.
Note
As the name implies, non-minimal phase systems have the same amplitude as
minimal phase equivalents but their phase angle is larger. To illustrate this consider
the Bode Diagram:
Ch 13: Alternative Frequency Response Plots 179
1 1
for q(s) = and compare this with the Bode Diagram of q(s) = . The
1+ s 1− s
magnitude is identical and the phase angle is less (minimal, in fact).
The NICHOLS CHART plots the open loop frequency response models on the
logarithmic axes for complex numbers:
and has its origin at (-180° , 0dB). The Nichols Chart origin defines the critical point
where q(jω) has a magnitude of 1 (i.e. 0dB) and a phase angle of -180°. It is
important for Gain and Phase margin evaluation.
As with Bode Diagrams, the Nichols Charts can be used to test for stability of the
closed loop system from a plot of the open loop transfer function, PROVIDED the
open loop system is stable. The stability test is simply that the closed loop system is
stable if the plot of q(jω) passes to the right of the origin, while it is unstable if q(jω)
passes to the left. These conditions are illustrated graphically in the following
Nichols Charts.
Ch 13: Alternative Frequency Response Plots 181
Recall that Phase Margin of the open loop system q(jω) gives an indication of the
damping factor that will be exhibited by the closed loop system h(jω), while the
Gain Margin gives an indication of its sensitivity to modelling errors. (Such errors
usually occur in the model g(jω) assumed for the process.)
Lag Circuit
The lag circuit increases low frequency gain and/or decreases high frequency gain
but introduces more phase lag. Its transfer function is
1 + sT
k(s) = K 1 < α ≤ 10
1 + sαT
Thus for K=1, the lag circuit moves the process transfer function g(jω) down on the
Nichols Chart at high frequencies (which is desired) but at the same time shifts it to
the left by an amount determined by α, especially at ωcm (which is bad).
Lead Circuit
The lead circuit introduces phase advance into the open loop which tends to stabilise
the closed loop system. Unfortunately it also amplifies high frequency noise. Its
transfer function is
1 + sαT
k(s) = K 1 < α ≤ 10
1 + sT
Thus with K=1 the lead circuit moves the process model, g(jω) or q(jω), to the right
(which is desired) but also moves it up at high frequencies (which is bad).
Recall that a good closed loop system has a gain margin greater than 2dB, a phase
margin in the range 30° to 70° and an Mmax in the range 1dB to 3dB. In the Nichols
Chart this defines the regions:
184 Control Engineering - 1
For closed loop stability and a reasonable phase margin (i.e. closed loop damping
factor) the open loop trace should lie to the right of the (0dB , -180°) origin.
As before, a lead circuit which swings to the right in the Nichols Chart would
achieve this. Setting α=4 gives a maximum phase advance of 37° at ωcm, without
introducing too much gain at high frequency. By inspection of the plot of g(jω) this
Ch 13: Alternative Frequency Response Plots 185
becomes
The compensated open loop system is minimal phase and passes to the right of the
(0dB , -180°) origin so the resulting closed loop system will be stable. Also the new
Phase Margin is 31° so the closed loop response should have a damping factor of
0.3 which is an improvement on the open loop response. Its Gain Margin is ∞ and
the gain could be changed to improve on the Phase Margin.
Once again stability of the typical closed loop system is determined by its
characteristic function
1 + f q(s)
N(s)
If q(s) = is a ratio of polynomials and f is a constant (independent of s) then
D(s)
function X(s) becomes
f N(s) + D(s)
X(s) =
N(s)
Clearly the zeros of X(s) are the poles of the closed loop system while the poles of
X(s) are the zeros of the open loop system.
As with the Nyquist diagram, stability of the closed loop system is determined by
traversing the Nyquist contour, D, in the s-plane, evaluating this function X(s) and
counting the number of likewise encirclements, <N>, of the origin of X(s).
where <Z> for this X(s) is the number of C/L unstable poles and
<P> for this X(s) is the number of O/L unstable zeros.
The Nyquist criterion then states that the closed loop system is stable iff <Z>=0, or
Ch 13: Alternative Frequency Response Plots 187
alternatively iff
< N >= − < P >= − p 0
The INVERSE NYQUIST PLOT is a plot of the inverse frequency response model
q ( jω ) on a polar plot. Hence the critical point (-f , 0) in the q - plane is the origin
of the function X(s) and the encirclements <N> are those that q ( jω ) make of the
critical (-f , 0) point.
Example
Determine the gains, f, that give a stable closed loop system for the open loop model
2
q(s) =
1 + 3s
as 's' moves around the Nyquist contour D gives the Inverse Nyquist Plot:
Since <P> = <No of unstable O/L zeros> = 0 here, stability of the closed loop
system requires that <N> = 0 which is true for all gains −f < 1 / 2 . Thus f > −0. 5
for stability (which includes all gains in a negative feedback configuration).
188 Control Engineering - 1
These are obtained in a similar manner to the Nyquist plot. For example:
1 1
Gain Margin = = 20 Log10 ( ) in [dB]
q of qof
One of the attractions of the Inverse Nyquist plot is the simple relationship that
exists between the inverse models of the open and the closed loop systems. For the
typical control loop:
h −1 (s) = f + q −1 (s)
Then an Inverse Nyquist plot of q (s) gives h(s) by a mere shift in the axis since:
Figure 13.26 Open and Closed Loop on the inverse Nyquist Plot
so
1
Magnitude[h(s)] =
M
Since points on the Inverse Nyquist Plot are q ( jω ) = x + jy the equation for the M
contours is simply
1
(1 + x)2 + y 2 = assuming that f=1
M2
Angle[h(jω)] = N
so that
Angle[h(s)] = −N
This gives straight lines radiating from the critical point at (-f , 0) in the Inverse
Nyquist plot at angles −N :
When designing closed loop systems in the Inverse Nyquist Plane, it is important to
know what lag/lead compensation does to the graphs.
Lag Circuit
Lead Circuit
The lead circuit improves on the phase margin for q(jω). It does however amplify
high frequency noise and should be used with caution. Its transfer function is
1 + sαT
k(s) = K 1 < α ≤ 10
1 + sT
g ( jω ) = −0. 25 ω 2
The Inverse Nyquist diagram for the uncompensated open loop system is:
Compensating the helicopter using a lead circuit with α=4 will ensure a phase
advance of φcm=37°. Setting T=0.25, gives this phase advance at a reasonable
frequency of ωcm=2 [rad/s].
The new Inverse Nyquist Plot for the compensated open loop system is:
Ch 13: Alternative Frequency Response Plots 193
Clearly its closed loop is stable as it does not encircle the critical point (-f , 0). Its
damping factor exceeds 0.3 (as the Phase Margin is over 30°) and is assumed
adequate. The overshoot to a step in the setpoint is expected to be about 45[%].
Control element k(s) affects the open loop model q(s) directly and linearly, since
q(s) = g(s) k(s)
and the closed loop model h(s) indirectly and non-linearly, since
= {1 + gk(s)}−1 gk(s)
The theoretical methods developed here all represent the open loop system model
q(s) in one form or another, with the express purpose of giving the designer insight
into the design problem. This allows systematic selection of optimal cascade
compensation elements, k(s). In addition the representations of q(s) often give
indirect indications (eg. M-contours) of the effects that the open loop system model
q(s) has on the closed loop system model h(s), thus aiding the design further.
Each design method has its own particular strengths and weaknesses, so it is
important in practice to select that method which relates most closely to the problem
in hand. Features of the various design methods are now summarized.
Ch 14: Summary of Linear Design Methods 195
Contours of closed loop poles in s-plane as open loop gain is varied from 0 to +∞.
Contours of closed loop poles in s-plane as some open loop parameter, other than
gain, is varied from 0 to +∞. (N.B. Rules for drawing root loci and characteristic
loci with gain variations from 0 to -∞ also exist.)
ω)
14.2. Frequency Response Methods, q(jω
14.3. Comment
(1) Not all control loops found in practice fit exactly into the form, like
unity feedback, required by specific design methods. Thus it is
encumbent on the engineer to modify the actual loop configuration for
analysis, if not implementation, until it fits that handled by the
proposed design method. Loop modification is easily achieved by
Block Diagram Algebra. Obviously the practical implications of such
alterations must be considered carefully.
(2) Alternative contours in the s-plane can be used to draw different Polar
plots for q(s) which may highlight other properties of the closed loop
system than merely stability. For example, in engineering applications
it is often desirable to contain the poles of the closed loop system
within the area of the s-plane shown shaded in:
Note
1 + sI + s2 DI
= K{ }
sI
Thus the PID controller puts a single pole at s=0 and two zeros at arbitrary positions
anywhere in the s-plane. The pole ensures Type 1 behaviour while the zeros attract
undesirable root loci towards the left in the s-plane.
1 + sI + s2 DI
k(s) = K{ }
sI(1 + sT)
Four versions of this popular control algorithm are commonly available in electronic
hardware. These are:
Note
(3) Should a non-zero error persist because the plant input has reached its
limits then the integral term grows without bound. Large values for
integral action cause the plant input to remain at its physical limit long
after the error has changed sign (indicating that a smaller plant input is
required). This induces excessive oscillatory behaviour in the loop that
can easily be avoided. Good PID controllers are designed to prevent
such reset wind-up in the integral term by limiting it to a value that
represents full-scale for the plant input u(s).
(4) The controller signals, <SP, PV and CV>, are often 4-20[mA] or
0-10[V], though many other possibilities exist. The requirements for a
particular application must be assessed to ensure that the correct
hardware is ordered.
The effect that a PID compensator has on a process model g(jω) can be estimated by
the following approximate reasoning based on a few critical points from the
frequency response model k(jω) for the PID controller. (Similarities between the PI
control and a Lag circuit plot at high frequencies, and between the PD control and a
Lead circuit plot at low frequencies should also be noted.)
Ch 15: Compensation Techniques 201
In industry the PID control algorithm is often installed in a loop and its parameters
found by trial-and-error adjustments (known as TUNING the controller). Such
methods for setting controller parameters appear adequate on the surface but give no
indication of how sensitive the closed loop system is to the inevitable changes in the
process model, g(s), or to measurement noise. Also, should the design call for a
complex compensator then a tuning procedure results in a sub-optimal design, and
in severe cases may never converge to a stable closed loop system.
General comment:
(1) The I-term is introduced to increase the low frequency gain of the
open loop system model q(s) and hence ensure that the closed loop
model h(s) tracks setpoints accurately (Type 1 control).
(2) The D-terms is intended to stabilise the loop by improving its phase
margin.
202 Control Engineering - 1
Example 1 PD Compensation
Notice from Fig 15.2 that a PD controller introduces a 45° phase advance at a
frequency of ω=1/D [rad/s]. Applied to the chemical plant this would improve the
phase margin of the combined open loop system q(s), and hence on the damping
factor of the closed loop system h(s).
To design the PD parameters, start by setting D=1 which would give a phase
advance of 45° at a frequency of ω=1 rad/s, and hence more phase advance at higher
frequencies (specifically at ω=2 rad/s where it is required). The PD compensator
transfer function is
k(s) = K(1+s) In theory
or
1+ s
k(s) = K In practice
1 + 0. 01s
Ch 15: Compensation Techniques 203
The Nyquist plot of the new q(s) is shown in the following figure with both the
theoretical and the practical k(s).
The Phase margin of 75° achieved by this PD compensation is above the value of
70° considered acceptable for mechanical systems, and well above the 30° value
recommended for many chemical processes. Since the D term is prone to noise
problems, it is advisable to use as small a value for D as possible. With this in mind,
the D-value could be reduced thereby minimising any noise problem while still
maintaining an adequate Phase Margin. In an actual design the parameter D would
be varied until a suitable compromise was reached between closed loop damping
factor (indicated by open loop phase margin) and noise amplification (caused by the
D term).
Example 2 PI Compensator
At steady state (ω=0) the open loop model q(j0) lies on the M=0.5 circle in the
Inverse Nyquist plot. Thus the closed loop system has a gain of h = 0. 5 and the
process output, y, at steady state is given by
y = 0.5 r
A PI controller can improve on this situation. Here the transfer function for the
compensator is
1 + sI
k(s) = K
sI
jω ) = jω I
k(
K(1 + jω I)
Start by setting I=1 [sec] (Since g(s) has a phase lag of 45° at ω=1 [rad/s] and a PI
controller with I=1[s] introduces another 45°, this gives a reasonable total phase lag
of 90° in the open loop model). The inverse Nyquist plot for q(s) becomes:
Ch 15: Compensation Techniques 205
From the plot of q(s)
= k(s) , it is clear that:
g(s)
Note that a slightly smaller value of I would cause q(jω) to bend to the left and so lie
closer to the M=1 circle for low frequencies, thus improving on the design.
(This does however have the disadvantage of increasing the sensitivity of the design
206 Control Engineering - 1
to changes in model g(s) and to external signals, d(t), disturbing the output, y(t) as
discussed in Chapter 16.)
In process control the first block g1(s) is usually faster than the second block g2(s).
The function of the minor loop is then to ensure quick elimination of any
disturbances to input u2(s).
Example
Here a tachometer measures speed and yields the transfer function model
1 Ω (s)
g1 (s) = =
B + sJ E(s)
The transfer function for the closed loop system (including the minor loop feedback)
is then
y(s) k
= h(s) = 2
v(s) s J + (B + f )s + k
Comparing this to the transfer function of the standard second order system
ω 2n
s + 2 ζω ns + ω 2n
2
it is obvious that the minor loop, f, directly affects the damping factor, ζ, and hence
characteristics of the closed loop response.
Thus minor loop is a very important configuration for motor position control since
damping factor, ζ, determines (closed loop) pole positions in the s-plane as follows:
208 Control Engineering - 1
u(s) k 2 (s)
=
v(s) 1 + k 2 (s)m(s)
Now, if m(s) is a model of the plant dynamics, g(s), then the transfer function of the
process (or open loop system) for the outer loop is
y(s) k 2 (s)
= g(s)
v(s) 1 + k 2 (s)m(s)
Thus the function of the inner loop is to cancel the plant dynamics (exactly in
theory, approximately in practice). Module, k1(s), then introduces any desired
characteristics for the open loop system of the outer loop to give the specified closed
loop performance.
It is strongly recommended in practice that every effort is made to reduce the dead
time term, τ, by altering the circuit itself, rather than to correct the problem by
feedback control. However, when this is not possible, the SMITH-PREDICTOR is
an ingenious scheme that alleviates the problem.
The transfer function m(s) is a model of the stable function g(s) that defines plant
dynamics without its dead-time. The output from Model 1 is intended to cancel the
process output exactly. The controller k(s) is designed (e.g. by Nyquist plots) to
control Model 2, the plant model m(s). This model does not contain a troublesome
dead-time term and design of k(s) is generally simple. Any discrepancies between
g(s) and m(s) are taken care of by the signal path through Model 1.
10(1 + s)
=
11 + s
10
= at low frequencies
11
which has a pole at s = −11 and is much faster than the original plant which had a
pole at s = −1 .
The transfer function, h(s), for the complete control system is then
y(s) 10
=
r(s) s2 + 11s + 10
which has poles at s = −0. 5 ± j0.87 . This closed loop system is slower than the
open loop plant, and slightly underdamped.
A major advantage of this method is that internal variables from the model are
readily available for use in minor loop controls for the inverse response loop. For
example, the model m(s) for a servo system could easily provide an estimate of the
speed of the motor and the inner inverse response loop could then incorporate
velocity and position feedback WITHOUT the need for a tachogenerator. (Such
212 Control Engineering - 1
designs, making full use of the process model, are studied in STATE SPACE
methods with particular reference to STATE OBSERVER systems).
e −2s
y(s) = u(s)
1+ s
and
r(s) − w(s)
u(s) =
s
1 e −2s e −2s
w(s) = u(s) + { − }u(s)
1+ s 1+ s 1+ s
1
= u(s) IFF the process is modelled exactly
1+ s
And so
1 1
u(s) = r(s) − u(s)
s s(1 + s)
Ch 15: Compensation Techniques 213
giving
s +1
u(s) = r(s)
s2 + s + 1
e −2s
y(s) = 2
r(s)
s + s +1
Without the models, the loop, fitted with the same controller, would be:
e −2s
q(s) =
s(1 + s)
Thus use of the Smith predictor allows a tighter control than would otherwise have
been possible.
Poles and zeros of k(s) are often chosen to be in the stable left-half s-plane and the
number of poles must equal (seldom) or exceed (generally) the number of zeros.
The real pole is known as the Roll Mode, while the oscillations are known as Dutch
Roll.
A suitable compensator k(s) for the aircraft deals with the two complex poles (the
Dutch roll mode) using two zeros in the same vicinity in the s-plane. Two poles then
need to be included in k(s) to make the compensator realisable. The compensator
probably also speeds up the dominant roll mode and eventually ends up being a
third or higher order transfer function.
Simple feedback control acts to reduce the effect of the disturbance since
q g
y(s) = r(s) + d d(s)
1+ q 1+ q
and a good design ensures that q(s) is very large (especially at low frequencies) so
that the second transfer function, between the disturbance and the plant output,
tends to zero.
Ch 15: Compensation Techniques 217
the transfer function, kd(s), for the feedforward controller is designed from the
equation
g d (s)
k d (s) =
g(s)
IFF this is realisable (i.e. stable and causal). Otherwise kd(s) should be designed to
approximate the transfer function ratio.
1 2s + 2s2
= r(s) + d(s)
1 + s + s2 2 + 3s + 3s2 + s3
The transfer function between d(s) and y(s) in the closed loop system has a Bode
plot:
which shows clearly that the closed loop system AMPLIFIES disturbances, d(s),
around a frequency of 1 rad/sec. In practice this means that the plant operator will
notice a deterioration in the performance of his plant under automatic control (and
will immediately condemn the entire control system).
For this system the closed loop performs according to the transfer functions
1
y(s) = r(s) + {≈ 0} d(s)
1 + s + s2
so the effect of the disturbance d(s) is eliminated (at least in theory). In practice
there will be some mismatch between the model used for the design of the
feedforward element and the plant. However feedforward is still likely to improve
on the performance of the overall system. Notice that the cost of the extra
feedforward control is an instrument that measures the disturbance d(s).
The block diagram of a typical industrial control problem includes components that
model these effects:
220 Control Engineering - 1
The controller required to deal with these problems comprises four modules:
16.1. Sensitivity
Mathematically, the sensitivity function, S(s), of a quantity, h(s), with respect to
variations in parameter, K, is defined by
h ∂h/h ∂ ln[h]
SK = =
∂ K / K ∂ ln[K]
Fractional change in h(s)
=
Fractional change in K
222 Control Engineering - 1
Sensitivity functions are combined by the chain rule. For example, the sensitivity of
h with respect to K is given by
h
SK = Sqh Sqg SgK
1
=
{1 + q(s)}2
The sensitivity function Sqh depends on the open loop model, q(s), only, so it is
possible to define a region in the q-plane or Nyquist diagram for which
Sqh > 1
indicating a sensitive design. This criterion defines a disc in the Nyquist plot, of
radius (1/f), centred on the critical point (-1/f , 0).
The formula for the disc is easily derived from the definition of Sqh since
1
Sqh = ≤β (where β is a set constant. E.g. β=0.2)
1 + q( jω )
becomes
1
(1 + x)2 + y 2 ≥ when q(jω) = x+jy
ß2
The mathematical boundary between sensitive and insensitive systems occurs when
ß=1. Graphically this defines the following circle (and disk) in the Nyquist plane.
Obviously good designs will attempt to avoid entering the disc. In practice this is
224 Control Engineering - 1
usually possible for low frequencies, but most designs cannot avoid the disc at high
frequencies.
To indicate regions in the Inverse Nyquist plot where the system is sensitive, the
sensitivity function
1
Sqh = ≤β
1 + q( jω )
can be re-written as
jω )
q( x + jy
= ≤β
jω )
1 + q( 1 + x + jy
when q ( jω ) = x + jy . Once again this defines a set of circles on the Inverse Nyquist
Plot. (These are similar to the M-circles found on Nyquist Plots).
In the Inverse Nyquist diagram this inequality for ß<1 defines the region to the right
of the straight line through the point (-f/2 , 0). Thus systems that lie to the right of
this line are robust (i.e. insensitive to modelling errors and disturbances).
Thus sensitivity of the closed loop system, h, to variations in the feedback element,
f, is given by the sensitivity function
fq(s)
Sfh (s) = −
1 + fq(s)
= -1 when fq(s) >> 1
This shows clearly that feedback does not reduce the sensitivity of a closed loop
system to changes in its feedback element f(s). Since f(s) models the measurement of
y(s), it is VITAL to note that the feedback configuration does not enhance the
accuracy of instrumentation (including any mathematical models in this path).
An electric motor is used to coil steel wire unto a drum. It has a transfer function
Ω(s) 50
= [rad/Volt-sec]
E(s) 1 + 7 s
where Ω(s) is the rotational velocity and E(s) is the input voltage to the motor.
As the steel wire is wound onto the drum, the process time constant changes from
100 milliseconds through 7 seconds to 500 seconds, because it is a function of the
system inertia. (Its nominal value T=7[s] is used extensively in the calculations.)
Using a cascade compensator k(s) = 10, sensitivity of the closed system, h(s), to
variations in the time constant of the open loop model, q(s), is given by
1
Sqh (s) =
1 + q(s)
1 + 7s
= ... (16.1)
501 + 7s
At steady state
226 Control Engineering - 1
1
Sqh ( j0) =
501
so the closed loop response will be minimally affect by changes in the open loop
model q(s), at very low frequencies.
Sqh ( j∞ ) = 1
Closed loop sensitivity to changes in the process time constant are computed from
the sensitivity function
50
g(s) = with 0.1 ≤ T ≤ 500 [s] and T=7[s] nominally
1 + sT
so
-50s
∂g= ∂T (by differentiation)
(1 + sT)2
and hence
∂g − sT ∂ T
=
g 1 + sT T
yielding
Combining Equations (16.1), (16.2) and (16.3), the sensitivity of h(s) to variations
in T is computed from the chain rule
1 + sT − sT
= 1
501 + sT 1 + sT
− sT
=
501 + sT
Hence the sensitivity of the closed loop system to changes in the time constant T is
−7 s
STh = for T=7[s], nominally
501 + 7s
Note
Figure 16.5 Bode Plots for the Closed Loop Transfer Function
Observe that the relative change in the time constant, ∂T/T, is 70. Thus the
sensitivity function STh must be very small (<1/70 or < -36.9[dB]) to ensure that the
relative change in the closed loop system ∂h/h is reasonable (< 1). This occurs for
frequencies below ω=1 [rad/s] as shown in the Bode diagram of the sensitivity
function.
If it is assumed that the setpoints in both loops are zero then the equations for the
original loop are
Ch 16: Sensitivity and Disturbance Rejection 229
u = k(s) [r − y] = − k(s) y
y = g(s) u + v
Hence
k(s)
−u = v The Process
1 + kg(s)
Also
v = ∆g(s) u The Controller
Then any design plot (e.g. Nyquist, Root Locus, etc.) of the process shows clearly
the range of controllers (i.e. parasitics, ∆g) that the designed loop can tolerate.
in a feedback configuration.
230 Control Engineering - 1
This factor is exactly the same as the function derived for loop sensitivity, namely,
1
Sqh (s) =
1 + fq(s)
In the Nyquist plot for open loop model q(jω), the region in which the disturbance is
amplified is easily identified by the same circle as before:
Disturbances are amplified for all frequencies within the shaded disc, so this region
should be avoided.
Example
Assuming that the disturbance is a simple sinusoid of unity amplitude and frequency
d(t) = 1.0 sin t
(1) Its Nyquist plot enters the sensitivity disk for frequencies above
0.7[rad/s], and
However for ω=0.2 the closed loop improves performance since a disturbance
d(t) = 1.00 sin 0.2t
Once again this information is obtained directly from the Nyquist plot during the
design.
is most useful. It defines the rate of change of pole position, s, with respect to the
parameter K and is used in Root Locus studies.
Example
Its parameters, A and B, were derived from a series of step tests performed on the
plant. The results were
A = 8.1 ± 2.6 [m]/[V] and B = 9.3 ± 1.2 [sec]
Using a PI controller
(1 + s)
k(s) = [V] / [m]
s
the nominal open loop transfer function, q(s), for the process model is
A(1 + s) 8.1(1 + s)
q(s) = = [V] / [V]
s(1 + sB) s(1 + 9.3s)
and the poles for the nominal closed loop system are at
sp = −0 . 489 ± j0.795 = sp (A, B)
Variations in Parameter A
Re-arranging this equation gives closed loop pole sensitivity with respect to A
234 Control Engineering - 1
∂s − (1 + s)
= SsA =
∂A 1 + A + 2Bs
− (1 + sp )
=
1 + A + 2Bsp
On the s-plane this uncertainty means that the pole position, p, is given by the range
p = s p ± ∆s p
This indicates that the closed loop system becomes less damped, with a lower
frequency of oscillation, as the open loop gain decreases. The exact extent of the
change can be quantified by the analysis. (Since the poles of a real system are
symmetric about the real axis in the s-plane the results follow immediately for the
other pole s p = −0 . 5 + j0.8 .)
Ch 16: Sensitivity and Disturbance Rejection 235
For interest, actual computation of the closed loop pole positions shows that
s p = −0 . 629 − j0.868 when ∆A = +2.6 (i.e. A = +32% change)
and
s p = −0 . 350 − j0. 685 when ∆A = -2.6 (i.e. A = -32% change)
Variations in Parameter B
Analysis to determine how changes in parameter B affect closed loop pole positions
is similar to the above. Thus only the main results are given below.
∂s − s2
=
∂B (1 + A + 2Bs)
= −0.053 − j0.027 when sp = −0 . 489 − j0. 795
= 0 . 071 ∠ − 153°
For interest, actual computation of the closed loop pole positions shows that
sp = −0 . 433 − j0.764 when ∆B = +1.2 (i.e. B = +13% change)
and
sp = −0 . 562 − j0.827 when ∆B = -1.2 (i.e. B = -13% change)
∆s p = SsA ∆A + SsB∆B
applies. The magnitude of changes in A and B would identify a region around the
nominal closed loop pole positions in the s-plane:
Ch 16: Sensitivity and Disturbance Rejection 237
These areas indicate the extent to which the closed loop system would be
downgraded by changes in the open loop model.
An Alternative Approach
Recall that the variation of parameters can also be dealt with by Root Locus
methods, as described earlier. Changes in A are nothing more than changes in the
Root Locus Gain, γ, while changes in B define a Characteristic Locus based on the
characteristic equation
1 9. 1
s2 + (A + {1 + A}s) = s2 + (s + 0.11)
B B
where γ is the Root Locus gain, and the numerator and denominator polynomials are
defined by:
238 Control Engineering - 1
m
N(s) = ∏ (s − z i ) Open loop zeros, zi
i=1
n
D(s) = ∏ (s − p i ) Open loop poles, pi
i=1
∂s − N(s)
=
∂γ γ ∂N +∂D
∂s ∂s
∂s 1 D(s)
=−
∂ pi s − pi γ ∂ N ∂D
+
∂s ∂s
∂s γ N(s)
=−
∂ zi s − zi γ ∂ N + ∂ D
∂s ∂s
Assuming that all signals remain within a range defined by the supply voltage, Vs ,
the operational amplifier circuit:
240 Control Engineering - 1
behaves in such a way that the output voltage is a negated sum of all input voltages,
individually scaled. (For clarity this circuit shows two inputs, e1 and e2, only.)
Mathematically this means that
R R R
e0 = −[ e1 + e2 + e3 + ... ]
R1 R2 R3
As many additional inputs can be included in the circuit as necessary (though large
systems must remain electronically sound.). Note that a control study defines only
constants, c1,c2,..., so the exact choice of resistor values, R, R1,R2,... depends on
electronic criteria (e.g. OpAmp input impedance, Leakage current, etc.).
acts as an integrating circuit in that the output voltage is a negated integral of the
sum of all the input voltages, individually scaled. (Once again a two-input circuit is
Ch 17: Electronic Circuitry from Transfer Functions 241
ò ò ò
1 1 1
e0 = − [ e1dt + e2 dt + e3dt + ...]
R1C R 2C R 3C
c1 c c
= −[ e1 + 2 e 2 + 3 e3 + ...]
s s s
Control theory provides the constant values c1,c2,..., so the exact choice of capacitor
value, C, and resistor values, R1,R2,... comes from electronic considerations.
This circuit is more difficult to deal with than the summation circuit since it includes
a capacitor that can retain a charge. Practical circuits must take this into
consideration. In control applications, the capacitors need to have correct initial
voltages on them to prevent unforeseen bumps to the system when the controller is
switched in. Thus a practical circuit contains numerous switches in addition to the
basic circuit shown above. In cheaper controllers an extremely large resistor is often
placed across the capacitor to leak off excess charge. This saves on switches, but
may cause a bumpy transfer of control that would not be tolerated on industrial
plants. In fact control theory and electronics are only the beginning of the design
and considerable effort has to go into commissioning the system to make it
acceptable to all concerned under operational conditions.
As it stands this diagram does not link directly with the above electronic circuits (i.e.
There is no one-to-one mapping between this diagram and OpAmp circuits).
However skilful use of block diagram algebra can convert the original comparator
circuit to a block diagram that is more in line with those for Summation and
Integration in OpAmp circuits . (It is not clear exactly what-to-do-when, so a little
experimentation is essential during this block-shuffling exercise. There are many
possibilities so the chances of success are good.)
This block diagram uses an OpAmp that sums two signals, +r and -y, to give -e (It
has been devised with extensive reference to the OpAmp summer circuit). To obtain
signal -y from +y and signal +e from -e, two single-input summers are needed. The
final block diagram result has exactly the same transfer characteristics as the original
comparator. Thus the electronic circuit requires three operational amplifiers, as
indicated, to mimic the comparator (providing all the signals remain within the
power rails.).
First cancel all common factors (i.e. Equal poles and zeros) between the numerator
and denominator polynomial. Then consider eliminating all factors in which poles
and zeros are nearly equal (Using the principle of Pole-Zero cancellation) and also
non-dominant poles (The principle of Pole dominance). Such preparation will
simplify the electronic circuit and may have little or no effect on the performance of
the final control system; but check that this is a valid assumption and ensure
causality (i.e. m ≤ n) in the final transfer function.
Consider
and assume that k(s) is in its simplest possible form (i.e. all common factors have
been cancelled, etc.).
244 Control Engineering - 1
(a) If m=n (as it does here) then divide the numerator by the denominator to give
a o + a1s a' 2
k(s) = +
b o + b1s + b 2s2 b2
or
a o + a1s u(s)
k(s) = +α =
b o + b1s + b 2s 2 e(s)
Note that the controller output, u(s), can be written as the sum of two terms
a o + a1s
u(s) = e(s) + α e(s) = u1 (s) + u 2 (s)
b o + b1s + b 2s2
To continue, concentrate on the first term alone, and execute the following
procedure (which converts k(s) from a transfer function to a state-space model, but
can be regarded as a recipe required to find OpAmps circuits. Although it appears
complicated, it does produce the simplest electronics most easily. Also, state space
theory is useful in manipulating the design further, if necessary).
(b) If m<n then multiply both denominator and numerator of k(s) by an internal
variable, w(s), as follows
u(s) a o + a1s w(s)
k(s) = = ∗
e(s) b o + b1s + b 2s2 w(s)
Note that there are exactly n (=2) equations here. These define the relationships
between two internal variables, [w(s)] and [sw(s)], and the transfer function input,
e(s). The second equation looks a bit odd but is very simply stating that the
derivative of variable [w(s)] is the variable [sw(s)] (Obviously). The brackets [ ] are
used to indicate an electronic signal, wherever this might be unclear. In this
example, signals [w(s)] and [sw(s)] are two important variables of the electronic
circuit and are known as state variables [x1(s)] and [x2(s)].
Ch 17: Electronic Circuitry from Transfer Functions 245
The electronic circuit is then constructed from a set of integrators (and invertors,
where necessary) since
ò
[sw(s)] = s[sw(s)]dt
and
ò
[w(s)] = s[w(s)]dt (mixing mathematical notations)
The resulting block diagram is readily developed directly from the above Equation
17.1. Start with the input [s2w(s)] to an integrating block:
Then integrate [s2w] to give [sw] and finally [w]. Finally, connect the signal lines:
To complete the transfer function collect the necessary terms, [sw] and [w] (and if
m=n, then [e] as well), to form the transfer function output, according to Equation
17.2. The required block diagram is:
246 Control Engineering - 1
In practical circuits it is extremely important to ensure that all the electronic signals,
especially [sw] and [w], remain within the range set by the OpAmp power rails. The
Final Value Theorem, Initial Value Theorem and digital simulation can assist in
finding extreme values for the circuit signals, using the transfer functions:
u(s) a o + a1s
=
e(s) b o + b1s + b 2s2
w(s) 1
=
e(s) b o + b1s + b 2s2
[sw(s)] s
=
e(s) b o + b1s + b 2s2
[s2 w(s)] s2
=
e(s) b o + b1s + b 2s2
and assuming an appropriate e(s). (Note that larger n values need more equations.)
If the signals are too big (or too small) then return to the original equations, 17.1
and 17.2, and scale the variables [w] and [sw] by appropriate constant factors.
Substitute these new variables into the equations to generate a second set of
equations
b1 α b0 α
s[ α sw(s)] = − [ α sw(s)] − [ β w(s)] + e(s)
b2 β b2 b2
β
s[β w(s)] = [ α sw(s)]
α
These equations are then converted to a block diagram and checked. Remember this
is a design exercise, so iterate as required until the perfect circuit emerges.
Sometimes the constants, ci, are enormous (e.g. 20x107) and the gains of individual
blocks within the circuit may need additional reduction. Simple block diagram
Ch 17: Electronic Circuitry from Transfer Functions 247
algebra is useful in distributing such gains amongst the system blocks. (The key to
finding a successful circuit is experimentation at the block diagram stage.)
A Numeric Example
to an OpAmp circuit with input, e(s), and output, u(s). Assume that Laplace variable
(s) is in [rad/s].
Sample Solution
(1) Note (a) that m(=1) < n(=2) so that k(s) is realisable, (b) there will be
n(=2) differential equations to implement so n(=2) integrators are needed,
and (c) the transfer function is in its simplest form already; there are no
cancellations nor dominant roots since:
(2) Convert transfer function k(s) to an equivalent state space model by the
Direct Programming method as follows
u(s) 3 + 15s w(s)
k(s) = = ∗
e(s) 6 + 4s + 2s w(s)
2
Note that the basic OpAmp circuits, consisting of two integration units
and one summation unit, are indicated on the diagram as thin-lined
rectangular blocks, but could have been omitted until the final stages of
the conversion exercise.
(4) Add the blocks defined by the numerator equation in terms of [sw] and
[w] to form the transfer function output from the internal variables, [sw]
and [w]. (Include signal [e] when m=n.) The resulting block diagram is:
Note that the circuit consists of two distinct operations. First input, e, is
mapped to internal variables (or states), [sw] and [w], by a set of
integrations. Secondly the states, [sw] and [w], are mapped to the output,
u, by a set of summations. No matter how complex k(s) becomes this is a
standard form; only the number of integrators and summers change. (By
the way, the states are not uniquely defined.)
(5) Study this block diagram closely and try to simplify it by block diagram
algebra to give the final four OpAmp diagram:
Ch 17: Electronic Circuitry from Transfer Functions 249
(6) Convert each block to an OpAmp circuit and connect its signals to form
the final electronic circuit corresponding to the transfer function k(s). The
resulting electronic circuit is:
The control study provides the ratios of resistors and capacitors. The
engineer chooses the final component values. In this design:
clearly indicating which values were selected by the user (on electronic
criteria) and which were dictated by the transfer function k(s).
(8) Obtain step response signals for use in commissioning the electronic
circuit. The relevant transfer functions for this design are:
250 Control Engineering - 1
u(s) 3 + 15s
=
e(s) 6 + 4s + 2s2
w(s) 1
=
e(s) 6 + 4s + 2s2
[sw(s)] s
=
e(s) 6 + 4s + 2s2
[s2 w(s)] s2
=
e(s) 6 + 4s + 2s2
Digital Simulation results provide signal shapes and amplitudes that can
be used to determine circuit adjustments that improve on its internal
signals by optimising their amplitudes (both [w(s)] and [sw(s)] are a bit
small, compared to [u(s)], and could be scaled up). Such traces are also
useful for hardware commissioning and acceptance testing.
Initial conditions
An analog simulator for the process can be built from OpAmps by modelling its
transfer function g(s). It should include non-zero inputs, U(t), and outputs, Y(t), to
test the effect of initial conditions on the performance of the controller design.
This electronic circuit combination of g(s) and k(s) forms a convenient and cost-
effective test-bed on which to experiment with the proposed control strategy and its
hardware implementation. Effects such as switch-over from Manual to Automatic
control, down-grading due to fault conditions, the influence of limited signal
amplitudes on overall performance, long-term drift, thermal stability, etc. all need to
be thoroughly investigated. In this way the final product, the electronic controller, is
ruggedised to survive in harsh industrial environments, and the high cost of on-line
experimentation is minimised.
18
Engineering Applications
Engineering design should aim for elegant simplicity. Thus it is good practice to
start any controller design project by ensuring that the model derived for its process
contains only essential information (without becoming simplistic). The attention
devoted to modelling has a direct impact on the complexity of the final electronic
circuit produced by control theory and is thus important to its design.
Skilful use of such simple dynamic models provides vast amounts of useful
engineering information about a control problem and its solution. Ultimately the
insight gained from these data could mean the difference between success and
failure of the controller design.
The problem is to design a controller, k(s), for the unity feedback configuration:
so that:
(2) its output, y(t), tracks the setpoint, r(t), exactly for step inputs,
(4) its input, u(t), remains in the range ±4 [V] for demanded setpoint
changes of ±1 [mA].
Design
To track step setpoints with zero error requires a Type Number of 1. Selecting an
appropriate controller form, the open loop transfer function model is
4
q(s) = g(s) k(s) = [mA] / [mA]
s(1 + 7s)
Decide to use Nyquist Plots as the design tool. (Since there is no dead time in q(s)
the Root Locus would also work.) The Nyquist Plot of q(jω) also shows a plot of the
plant model g(jω) for comparison.
From the plot of q(jω) it is possible to predict (roughly, but adequately) the
dynamics of the closed loop system (i.e. Type number, Stability, Likely damping
factor, Closed loop bandwidth, etc.). Specifically it is observed that:
(2) The closed loop system will be stable (since N=0 and p0=0).
(3) The phase margin for q(jω) is small so the closed loop system will be
oscillatory with
The Final Value Theorem shows that the steady state error for step inputs, r=1/s, is
zero
s(1 + 7s) 1
e ∞ = lim e(t) = lim se(s) = lim s =0
t →∞ s→ 0 s→ 0 s(1 + 7s) + 4 s
while the Initial Value theorem shows that the initial input is also zero.
So the closed loop system, with this controller in place, tracks any constant setpoint
exactly. Digital simulation of g(s) and k(s)=1/s produced the following closed loop
step response data, confirming the predictions made from the Nyquist plot about its
oscillatory behaviour with a period of 9[s].
Ch 18: Engineering Applications 255
Note that the improved setpoint tracking has been achieved at the cost of stability,
since the closed loop system is much more oscillatory than the open process.
With a bias towards simplicity, it is possible to add either a pole or a zero to k(s). A
zero tends to swing q(jω) in a counter-clockwise direction. This is good since it will
increase the phase margin (and hence the damping factor). From the Nyquist plot it
is observed that the frequency closest to the critical point (-1, j0) is 0.7[rad/s], so a
zero at s=-0.7[rad/s] is a likely choice to start with. (Remember that experimentation
is a critical component of any design exercise.) Thus try adding the factor (1+2s) to
k(s), giving k(s) = (1+2s)/s where 2 is chosen as it is a round number near 1.4[s]
(which was itself computed from 1/0.7).
The resulting Nyquist plot, showing both the previous and the new q(jω), is then:
The observation that can be made from the new design are that:
These predictions are confirmed by the corresponding closed loop step response:
Engineering Conclusions
k(s) = K(1+sB)/s
about (near) the design values (K=1[°]/[°/s], B=2[s]). In each case deduce important
engineering criteria, such as:
For example, it is obvious that an increase in B decreases the overshoot, but requires
a simultaneous decrease in K to preserve the response time and initial input value.
The selection of K=1 and B=2 gives a reasonable design (but the final choice is
subjective, and K=2 with B=2 is also acceptable).
Should this simple artefact fail to produce an adequate solution then specialised
non-linear control engineering methods must be consulted and applied, or the
process itself should be altered (which in itself is sometimes a very useful
conclusion resulting from an application of control theory to a practical problem).
A robot arm joint driven directly from a motor shaft and operated in a gravitational
field is such an example. It is discussed here because it has many features that
illustrate the subtle difference between an engineering approach (that, in spite of the
risks, might work well and save design time) and a more fundamental scientific
258 Control Engineering - 1
investigation (with detail that, as it happens in this case, is not necessary to solve the
control problem).
Clearly the robot arm is inherently stable in the position shown in Figure 18.6a and
unstable in its normal operating position (Figure 18.6b). Thus it needs feedback
control to produce an acceptable final system.
The aim of the controller design is to hold the robot arm at any position within its
operating circle by manipulating voltage to its motor field coils. Its angular position
is measure by a potentiometer.
The model relating <field coil voltage>, u(t), and <joint angle>, y(t), is based on a
block diagram that is derived from elementary laws of physics:
Ch 18: Engineering Applications 259
In the diagram, the field voltage produces a torque (the K-block). This must
overcome the effects of gravity (the N-block) and possibly drag (the B-block). Any
excess torque accelerates the arm (the J-block), producing angular velocity, Ω, and
hence angular position, θ, measured by a potentiometer (the A-block).
where L is the distance from the arm joint to the centre of gravity of the entire arm
system (as shown in Figure 18.6b), M is the equivalent mass of the robot arm and g
is the gravitational constant 9.8 [m/s2].
λ ( θ0 ) = Mg L cos ( θ0 )
System Identification
From the block diagram analysis of the process it was noted that
λ ( θ 0 ± π ) = − MgL cos ( θ0 ) = − λ ( θ0 )
Thus the model parameters (J, B, λ) derived for the robot arm in its open loop stable
position (below the horizontal) immediately give the parameters in its unstable
operating region (above the horizontal).
260 Control Engineering - 1
(Note that upward responses are different to downward ones the arm is highly
non-linear.) From these graphs a set of linear dynamic models were derived for the
robot arm in its stable position of operation. The nominal model (showing standard
deviations as well) for the robot arm in its stable operating region is
(0. 093 ± 0. 010)
g(s) = ... (18.1)
1 + ( 0 . 076 ± 0 . 013 )s + (0. 0111 ± 0. 0012)s2
Root Loci for the two linearized cases, Equations (18.1) and (18.2), predict similar
Ch 18: Engineering Applications 261
From these graphs it is concluded that a high gain proportional controller would
stabilise the robot arm both above and below the horizontal. Also, for positions
above the horizontal, the gain must exceed a minimum to ensure stable behaviour.
Note however that the linear design is not yet complete since the control system can
be much improved by feedback compensation in the form of a simple lead circuit.
The aim is to increase the damping of the controlled system. After a few attempts, in
which designs for g(s) in Equation (18.2) were tested on g(s) in Equation (18.1), the
final design for a robot arm feedback compensator was
50 + 4s
k(s) = ... (18.3)
1 + 0.02s
Sections of the Root Loci for q(s), with gains around the design value, are now more
damped in both linearized cases Also the differences between the Loci for the stable
and the unstable linear models is no longer very significant:
262 Control Engineering - 1
Controller Implementation
When implemented on the robot arm system, the closed loop results are impressive
in that its behaviour is extremely consistent, especially compared to the uncontrolled
robot:
The controlled robot is also very robust, as shown by its reaction to a variety of
setpoint changes and impulse disturbances during normal operation:
264 Control Engineering - 1
Conclusion
Bibliography
Aseltine, J.A. Transform method in linear system analysis. McGraw-Hill, New
York, 1958.
Bissell, C.C. Control engineering. Van Nostrand Reinhold, London, 1988.
Bolton, W. Control engineering. Longman, Harlow, 1992.
Borrie, J.A. Modern control systems. Prentice-Hall, Englewood Cliffs, 1986.
Brogan, W.L. Modern Control Theory. Prentice-Hall, Englewood Cliffs, 1985.
Chen, C-T. System and signal analysis. Holt, Rinehart & Winston, New York, 1989.
Chen, C-T. Linear system theory and design. Holt, Rinehart & Winston, New York,
1984.
D'Azzo, J.J. & Houpis, C.H. Linear control system analysis and design. McGraw-
Hill, New York, 1988. (3rd Edition)
Dorf, R.C. Modern Control Systems. Addison-Wesley, Reading, 1992. (6th Edition)
Franklin, G.F, Powell, J.D. & Emami-Naeini, A. Feedback control of dynamic
systems. Addison-Wesley, Reading, 1986.
Hostetter, G.H., Savant, C.J. & Stefani, R.T. Design of feedback control systems.
Holt, Rinehart & Winston, New York, 1982.
Karni, S. & Byatt, W.J. Mathematical methods in continuous and discrete systems.
Holt, Rinehart & Winston, New York, 1982
Kuo, B.C. Automatic control systems. Prentice-Hall, Englewood Cliffs, 1982. (4th
Edition)
Maciejowski, M. Multivariable feedback design. Addison-Wesley, Wokingham,
1989.
Miron, D.B. Design of feedback control systems. Harcourt Brace Jovanovich, San
Diego, 1989.
Morari, M. & Zafiriou, E. Robust process control. Prentice-Hall, Englewood Cliffs,
1989.
Nise, N.S. Control systems engineering. Benjamin Cummings. Redwood City, 1992.
Phillips, C.L. & Harbor, R.D. Feedback Control Systems. Prentice-Hall, Englewood
Cliffs, 1991. (2nd Edition)
Raven, F.H. Automatic Control Engineering. McGraw-Hill, New York, 1987. (4th
Edition)
Saucedo, R. & Schiring, E.E. Introduction to continuous and digital control
systems. Macmillan, New York, 1968.
Sinha, N.K. Control systems. Holt, Rinehart & Winston, New York, 1986.
Slotine, J-J.E. & Li, W. Applied nonlinear control. Prentice-Hall, Englewood Cliffs,
1991.
Van De Vegte, J. Feedback Control Systems. Prentice-Hall, Englewood Cliffs,
1986.
270 Control Engineering - 1
Notes
271
Notes (Continued)
272 Control Engineering - 1
Many of the design diagrams and simulation plots shown in these notes were
produced by sCAD, a computer-aided design program written by the author.
The original program was written in Pascal for DOS and has subsequently been
converted to Visual Basic running on Windows 98 and Windows NT. The latest
version of the program is available for download from the website:
www.ci.ee.uct.ac.za
Other computer aided design programs, zCAD and ppCAD, are also available
from this site. All these programs are continuously being upgraded in reaction to
student comment, and include comprehensive on-line help files.
Property of
Control & Instrumentation Laboratory,
Department of Electrical Engineering,
University of Cape Town.
Should you wish to continue using this product then please enquire about a licence.