2019handout1 Automatic Control Systems

Download as pdf or txt
Download as pdf or txt
You are on page 1of 91

AUTOMATIC CONTROL SYSTEMS

1. Introduction to Control Systems


1.1 Definitions

A System is an arrangement of physical components connected or related in such a


manner as to form and/or act as an entire unit.

A Control System is an arrangement of physical components connected or related


in such a manner as to command, direct, or regulate itself or another system.

In engineering the term control system is restricted to those systems whose major
function is to dynamically or actively command, direct, or regulate.

Example:

The figure shown below is an example of a control system which consists of a mirror
pivoted at one end and adjusted up and down with a screw at the other end. The
angle of reflected light is regulated by means of the screw.

The input is the stimulus or excitation applied to a control system from an external
energy source, usually in order to produce a specified response from the control
system.

The output is the actual response obtained from a control system. It may or may
not be equal to the specified response implied by the input.

input / stimulus output / response


Control
desired response system actual response

The purpose of the control system usually identifies or defines the output and input.
Given the output and input, it is possible to identify or define the nature of the
system’s components.

author w chiyabu Page 1


In identifying a system, spurious inputs producing undesirable outputs are not
normally considered as inputs and outputs in the system description.

A control system consists of subsystems and processes (or plants) assembled for
the purpose of obtaining a desired output with desired performance, given a
specified input.

1.2 Types of Control Systems

1. Man-made Control Systems


An electric switch is a man-made control system, controlling the flow of
electricity. By definition, the apparatus or person flipping the switch is not a part
of this control system.
Flipping the switch on or off may be considered as the input. The output is the
flow or non-flow of electricity.

2. Natural, including Biological Control Systems


The act of pointing at an object with a finger requires a biological control
system consisting chiefly of the eyes, the arm, hand and finger, and the brain of a
person.
The input is the precise direction of the object (moving or not) with respect to
some reference, and the output is the actual pointed direction with respect to the
same reference.

3. Control Systems whose components are both Man-made and Natural


The control system consisting of a person driving road motor vehicle has
components which are both man-made and biological.
The driver wants to keep the road motor vehicle in the appropriate lane of the
road way. He/She accomplishes this by constantly watching the direction of the
road motor vehicle with respect to the direction of the road. In this case, the
direction or heading of the road, represented by painted guide line or lines on
either side of the lane may be considered as the input. The heading of the road
motor vehicle is the output of the system. The driver controls this output by
constantly measuring it with his/her eyes and brain, and correcting it with his/her
hands on the steering wheel.
The major components of this control system are the driver’s hands, eyes and
brain, and the vehicle.

Reasons for Building Control Systems:


1. Power amplification
2. Remote control
3. Convenience of input form
4. Compensation for disturbances

author w chiyabu Page 2


1.3 Classification of Control Systems

Control Systems are configured or classified as either open-loop control systems or


closed-loop control systems. The distinction is determined by the control action,
which is that quantity responsible for activating the system to produce the output.
In order to classify a control system as open-loop or closed-loop, the components of
the system must be clearly distinguished from components that interact with, but
are not part of the system.

disturbance 1 disturbance 2
input + + output
input controller process
or or
transducer or plant
reference + + controlled
summing summing variable
junction junction

Open-Loop Control System

disturbance 1 disturbance 2
input + + output
or input + controller + process + or
reference transducer or plant controlled
variable

output
transducer
or sensor

Closed-Loop Control System

Open-Loop Control Systems


An open-loop control system is one in which the control action is independent of the
output because it has no feedback. In any open-loop control system the output is
not compared with the reference input. In other words, there is no comparison of
the controlled variable with the desired input. Each setting of the input determines a
fixed operating position for the control elements.
An open-loop system starts with a subsystem called an input transducer, which
converts the form of the input to that used by the controller. The controller drives a
process or a plant. The input is sometimes called the reference, while the output can
be called the controlled variable. Other signals, such as disturbances, are shown
added to the controller and process outputs via summing junctions, which yield the
algebraic sum of their input signals using associated signs.
The distinguishing characteristic of an open-loop system is that it cannot
compensate for any disturbances that add to the controller’s driving signal.

author w chiyabu Page 3


Advantages of Open-Loop Control Systems
1. Simple construction and ease of maintenance
2. Less expensive in comparison with corresponding closed-loop systems
3. There is no stability problem
4. Convenient when output is hard to measure or measuring the output precisely is
economically not feasible.

Disadvantages of Open-Loop Control Systems


1. Disturbances and changes in calibration cause errors, and the output may be
different from what is desired.
2. To maintain the required quality in the output, recalibration is necessary from
time to time.

Closed-Loop Control Systems


A closed-loop control system is one in which the control action is somehow
dependent on the output. A closed-loop (or feedback) control system is a
control system that tends to maintain a prescribed relationship of one system
variable to another by comparing functions of these variables and using the
difference as a means of control.
The input transducer converts the form of the input to the form used by the
controller. An output transducer, or sensor, measures the output response and
converts it into the form used by the controller.
The first summing junction algebraically adds the signal from the input to the signal
from the output, which arrives via the feedback path - the return path from the
output to the summing junction.
The result is generally called the actuating signal. However, in systems where
both the input and output transducers have unity gain (that is, the transducer
amplifies its input by 1), the actuating signal’s value is equal to the actual difference
between the input and the output.
Under this condition, the actuating signal is called the error.
The closed-loop system compensates for disturbances by measuring the output
response, feeding that measurement back through a feedback path, and comparing
that response to the input at the summing junction. If there is any difference
between the two responses, the system drives the plant, via the actuating signal, to
make a correction. If there is no difference, the system does not drive the plant,
since the plant’s response is already the desired response.
Closed-loop systems, then, have the obvious advantage of greater accuracy than
open-loop systems. They are less sensitive to noise, disturbances, and changes in
the environment. Transient response and steady-state error can be controlled more
conveniently and with greater flexibility in closed-loop systems, often by a simple
adjustment of gain (amplification) in the loop and sometimes by redesigning the
controller. We refer to the redesign as compensating the system and to the
resulting hardware as a compensator. On the other hand, closed-loop systems are
more complex and expensive than open-loop systems.

author w chiyabu Page 4


In summary, systems that perform measurement and correction are called closed-
loop, or feedback control, systems. Systems that do not have this property of
measurement and correction are called open-loop systems.
For systems in which the inputs are known ahead of time and in which there are no
disturbances it is advisable to use open-loop control systems. Closed-loop control
systems have advantages only when unpredictable disturbances and/or
unpredictable variations in system components are present.

Computer-Controlled Systems
In many modern systems, the controller (or compensator) is a digital computer. The
advantage of using a computer is that many loops can be controlled or compensated
by the same computer through time sharing. Furthermore, any adjustments of the
compensator parameters required to yield a desired response can be made by
changes in software rather than hardware. The computer can also perform
supervisory functions, such as scheduling many required applications.

1.4 Feedback

Feedback is that property of a closed-loop control system which permits the output
(or some other controlled variable of the system) to be compared with the input to
the system (or an output to some other internally situated component or subsystem
of the system) so that the appropriate control action may be formed as some
function of the output and input.
For a system to be classified as a feedback control system, it is necessary that the
controlled variable be fed back and compared with the reference input. In addition,
the resulting error signal must actuate the control elements to change the output
so as to minimise the error.

Characteristics of Feedback
1. Increased accuracy. For example, the ability to faithfully reproduce the input.
2. Reduced sensitivity of the ratio of output to input to variations in system
characteristics.
3. Reduced effects of non-linearities and distortions.
4. Increased bandwidth. The bandwidth of a system is that range of frequencies
(of the input) over which the system will respond satisfactorily.
5. Tendency towards oscillation or instability.

1.5 The Control Systems Engineering Problem


The nature of control systems engineering is the consideration of two problems: the
analysis and the design of a control system configuration.
Analysis is the investigation of the properties of an existing system. The design
problem is the choice and arrangement of control systems components to perform a
specific task.
Design methods are either by analysis or by synthesis.

author w chiyabu Page 5


Design by Analysis is accomplished by modifying the characteristics of an existing
or standard system configuration. Design by synthesis is accomplished by defining
the form of the system directly from its specifications.

Representation of the Problem: the Model


In order to solve a systems problem, the specifications or description of the system
configuration and its components must be put into a form suitable for analysis,
design, and evaluation.
In the study of control systems, the following basic representations (models) of
physical components and systems are extensively employed:

1. Differential Equations and other mathematical relations


2. Block Diagrams
3. Signal Flow Graphs

Block Diagrams and Signal Flow Graphs are shorthand, graphical representations of
either the schematic diagram of a physical system, or the set of mathematical
equations characterising its parts.
Mathematical models, in the form of system equations, are employed when detailed
relationships are required. Every control system may theoretically be characterised
by mathematical equations. The solution to these equations represents the system’s
behaviour.

1.6 Control System Terminology

Block Diagrams: fundamentals


A Block Diagram is shorthand, pictorial representation of the cause and effect
relationship between the input and output of a physical system. It provides a
convenient and useful method for characterising the functional relationships among
the various components of a control system. System components are alternatively
called elements of the system.

block
input output

The interior of the rectangle representing the block usually contains a description of
or the name of the element, or the symbol for the mathematical operation to be
performed on the input to yield the output. The arrows represent the direction of
unilateral information or signal flow.

Example:

input control output x d y = dx


element dt dt

author w chiyabu Page 6


The operations of addition and subtraction have a special representation.

Examples of summing points z


+
x + x+y x + x-y x + x+y+z

+ - +
y y y

Any number of inputs may enter a summing point. The output is the algebraic sum
of the inputs.

Some authors put a cross. Sometimes it is confused with multiplication operation.

In order to employ the same signal or variable as an input to more than one block or
summing point, a take-off point is used. This permits the signal to proceed
unaltered along several paths to several destinations.

x x take-off point
x x x
take-off x
point x

Block Diagram of a Feedback Control System


The blocks representing the various components of a control system are connected
in a fashion which characterises their functional relationship within the system.

disturbance
actuating manipulated u
reference signal variable controlled
input control plant output
r + e=r±b elements m g2 c
± g1
b
forward path
primary
feedback
signal
feedback
elements
h

feedback path

The arrows of the closed-loop connecting one block with another represent the
direction of flow of control energy or information, and not the main source of
energy for the system.

author w chiyabu Page 7


Generalised Feedback Control System

Terminology of the Closed-Loop Block Diagram


Lower case letters are used to represent the input and output variables of each
element as well as the symbols for the blocks g1, g2 and h. These quantities
represent functions of time, unless otherwise specified.

Example: r = r(t)

Capital letters denote Laplace transformed quantities as functions of the complex


variable s, or Fourier transformed quantities (frequency functions) as functions of
the pure imaginary variable jw. Functions of s are usually abbreviated to the capital
letter appearing alone. Frequency functions are never abbreviated.

Example: R(s) is abbreviated to R. R(jw) is never abbreviated.

1. The plant g2, also called the controlled system, is the body, process, or
machine, of which a particular quantity or condition is to be controlled.

2. The control elements g1, also called the controller, are the components
required to generate the appropriate control signal m applied to the plant.

3. The feedback elements h are the components required to establish the


functional relationship between the primary feedback signal b and the controlled
output c.

4. The reference input r is an external signal applied to a feedback control system


in order to command a specified action of the plant. It often represents ideal
output behaviour.

5. The controlled output c is the quantity or condition of the plant which is


controlled.

author w chiyabu Page 8


6. The primary feedback signal b is a signal which is a function of the controlled
output c, and which is algebraically summed with the reference input r to obtain
the actuating signal e.

7. The actuating signal e, also called the error or control action, is the
algebraic sum consisting of the reference input r plus or minus (usually minus)
the primary feedback b.

8. The manipulated variable m (control signal) is that quantity or condition which


the control elements g1 apply to the plant g2.

9. A disturbance u is an undesired input signal which affects the value of the


controlled output c. It may enter the plant by summation with m, or via an
intermediate point.

10. The forward path is the transmission path from the actuating signal e to the
controlled output c.
11. The feedback path is the transmission path from the controlled output c to
primary feedback signal b.
Negative feedback means that the summing point is a subtractor or e = r – b.
Positive feedback means that the summing point is an adder or e = r + b.

12. A Unity Feedback System is a feedback system in which the primary feedback
b is identically equal to the controlled output c.

Supplementary Terminology

13. A transducer is a device which converts one energy form into another.
Example: A potentiometer converts mechanical position into an electrical
voltage.

+
arm
v position position voltage
reference input input potentiometer output
voltage + v r
source r voltage
output
+

Schematic Block Diagram

14. The command v is an input signal, usually equal to the reference input r. But
when the energy form of the command v is not the same as that of the primary
feedback b, a transducer is required between the command v and the reference
input r.

author w chiyabu Page 9


command reference + e
input
transducer input
transducer
v r v r
-
b
c
feedback
transducer

15. When the feedback element consists of a transducer, and a transducer is


required at the input, that part of the control system is called the error
detector.

16. A stimulus or test input is any externally introduced input signal affecting the
controlled output c.
For example, the reference input r and a disturbance u are stimuli.
17. The time response of a system or element is the output as a function of time,
following the application of a prescribed input under specified operating
conditions.

Servomechanisms
A servomechanism is power-amplifying feedback control system in which the
controlled variable c is mechanical position, or a time derivative of position such as
velocity or acceleration.

Example:
A motor vehicle power-steering apparatus is a servomechanism.

control elements : g1 plant

gear + control hydraulic drive u wheels c


ratio r e valve amplifier linkage
angular - angular
position of the b position of
starting wheel road wheels
feedback
linkage
h

The command input is the angular position of the steering wheel. A small rotational
torque applied to the steering wheel is amplified hydraulically, resulting in a force
adequate to modify the output, the angular position of the front wheels. Negative
feedback is necessary in order to return the control valve to the neutral position,
reducing the torque from the hydraulic amplifier to zero when the desired wheel
position has been achieved.

author w chiyabu Page 10


Regulators
A regulator or regulating system is a feedback control system in which the
reference input or command is constant for long periods of time, often for the entire
time interval during which the system is operational.

A regulator differs from a servomechanism in that the primary function of a regulator


is usually to maintain a constant controlled output, while that of a servomechanism
is most often to cause the output of the system to follow a varying input.

2. Laplace Transforms
Laplace Transforms make solving linear Ordinary Differential Equations (ODEs) and
related initial value problems, as well as systems of linear ODEs, much easier.
Applications abound: electrical networks, springs, mixing problems, signal
processing, and other areas of engineering and physics.

Process of solving an ODE using the Laplace Transform Method:


Step1. The given ODE is transformed into an algebraic equation, called the
subsidiary equation
Step2. The subsidiary equation is solved by purely algebraic manipulations
Step3. The solution in Step 2 is transformed back, resulting in the solution of the
given problem

IVP AP Solving Solution


Initial Value Algebraic AP of the
Problem 1 Problem 2 by Algebra 3 IVP

The key motivation for learning about Laplace Transforms is that the process of
solving an ODE is simplified to an algebraic problem (and transformations). This type
of mathematics that converts problems of calculus to algebraic problems is known as
operational calculus.

In the Laplace Transform Method, transformation of the terms of the differential


equation yields an algebraic equation in another variable s. Thereafter, the solution
of the differential equation is effected by simple algebraic manipulation in the s
domain (the new variable is s rather than time t). To obtain the desired time
solution, it is necessary to invert the transform of the solution from the s domain
back to the time domain.

If f(t) is a function defined for all t ≥ 0, its Laplace Transform is the integral of
f(t) times e-st from t = 0 to ∞. It is a function of s, say, F(s), and is denoted by L(f);
thus

author w chiyabu Page 11


The Inverse Laplace Transform of F(s) is denoted by L-1(F).
Thus f(t) = L-1(F)
Note: L-1(L(f)) = f and L(L-1(F)) = F

Notation
Original functions depend on t and their transforms on s. Original functions are
denoted by lowercase letters and their transforms by the same letters in
capital, so that F(s) denotes the transform of f(t), and Y(s) denotes the transform
of y(t), and so on.

Laplace Transform: General Formulas

author w chiyabu Page 12


Table of Laplace Transforms

author w chiyabu Page 13


Table of Laplace Transforms (continued)

author w chiyabu Page 14


Table of Laplace Transforms (continued)

author w chiyabu Page 15


3. Transfer Functions
In control theory, transfer functions are commonly used to characterise the input-
output relationships of components or systems that can be described by linear, time-
invariant, differential equations.

The Transfer Function of a linear, time-invariant, differential equation system is


defined as the ratio of the Laplace Transform of the output (response function) to
the Laplace Transform of the input (driving function) under the assumption that all
initial conditions are zero.

The Transfer Function will allow separation of the input, system, and output into
three separate and distinct parts. The Transfer Function will also allow us to
algebraically combine mathematical representations of subsystems to yield a total
system representation.

Let us begin by writing a general nth-order, linear, time-invariant differential


equation,

where c(t) is the output, r(t) is the input, and the ai’s, bi’s, and the form of the
differential equation represent the system.

Taking the Laplace Transform of both sides,

This equation is a purely algebraic expression. If we assume that all initial conditions
are zero, the equation reduces to

(ansn + an-1sn-1 + ... + a0)C(s) = (bmsm + bm-1sm-1 + ... + m0)R(s)

Now form the ratio of the output transform, C(s), divided by the input transform,
R(s):

author w chiyabu Page 16


Notice that the equation above separates the output, C(s), the input, R(s), and the
system, which is the ratio of polynomials in s on the right. We call this ratio, G(s),
the transfer function and evaluate it with zero initial conditions.

The transfer function can be represented as a block diagram with the input on the
left, the output on the right, and the system transfer function inside the block.

Notice that the denominator of the transfer function is identical with the
characteristic polynomial of the differential equation. Also, we can find the output,
C(s) by using
C(s) = G(s)R(s)

Properties of Transfer Functions


1. A Transfer Function is defined only for a linear system, and strictly, only for time-
invariant systems.
2. A Transfer Function between an input variable and an output variable of a system
is defined as the ratio of the Laplace Transform of the output to the Laplace
Transform of the input.
3. All initial conditions of the system are assumed to be zero.
4. A Transfer Function is independent of input excitation.

Example: Find the transfer function represented by

Solution:

Taking the Laplace Transform of both sides, assuming zero initial conditions we have

sC(s) + 2C(s) = R(s)

The transfer function, G(s), is G(s) = C(s) / R(s) = 1 / (s + 2)

Electrical Network Transfer Functions

The Transfer Function can be applied to the mathematical modelling of electric


circuits including passive networks and operational amplifier circuits.

Equivalent circuits for the electric networks may consist of passive linear
components: resistors, capacitors, and inductors.

author w chiyabu Page 17


Voltage-current, voltage-charge, and impedance relationships for capacitors, resistors, and inductors

Note: v(t) = V (volts), i(t) in A (amps), q(t) in Q (coulombs), C in F (farads), R in Ω (ohms),


G in 1/Ω (mhos), L in H (henries)

In the analysis we combine electrical components into circuits, decide on the input
and output, and find the transfer function. Our guiding principles are Kirchhoff’s
laws. We sum voltages around loops or sum currents at nodes, depending on which
technique involves the least effort in algebraic manipulation, and then equate the
result to zero. From these relationships we can write the differential equations for
the circuit. Then we can take the Laplace Transforms of the differential equations
and finally solve for the Transfer Function.

Example: Transfer Function – single loop via the Differential Equation

Problem: Find the transfer function relating the capacitor voltage, VC(s), to the input
voltage, V(s) in the figure below.

RLC network

Solution
In any problem, the designer must first decide what the input and output should be.
From the question, we are to treat the capacitor voltage as the output and the
applied voltage as the input.
Summing the voltages around the loop, assuming zero initial conditions, yields the
integro-differential equation for this network as

Changing variables from current to charge using i(t) = dq(t) = dt yields

author w chiyabu Page 18


From the voltage-charge relationship for a capacitor, q(t) = CvC(t)
Substituting for q(t) yields

Taking the Laplace Transform assuming zero initial conditions, rearranging terms,
and simplifying yields
(LCs2 + RCs + 1)VC(s) = V(s)

Solving for the transfer function, VC(s) = V(s), we obtain

The RLC electrical network has been transformed into

Block diagram of series RLC electrical network

Step-by-Step General Solution


1. Take the Laplace transform of the equations according to the circuit elements
assuming zero initial conditions

For the capacitor V(s) = I(s) / Cs

For the resistor V(s) = RI(s)

For the inductor V(s) = LsI(s)

2. Now the transfer function is the impedance V(s) / I(s) = Z(s)

From [sum of impedances] I(s) = [sum of applied voltages]

Z(s) = Ls + R + 1/Cs

3. Instead of writing the differential equation first and then taking the Laplace
Transform, we can draw the transformed circuit and obtain the Laplace transform
of the differential equation simply by applying Kirchhoff’s voltage law to the
transformed circuit.

author w chiyabu Page 19


Laplace-transformed network

Summary of Steps
1. Redraw the original network showing all time variables, such as v(t), i(t), and
vC(t), as Laplace transforms V(s), I(s), and VC(s), respectively.
2. Replace the component values with their impedance values. This replacement is
similar to the case of dc circuits, where we represent resistors with their
resistance values.

Transfer Function for Series Circuit

General Series Circuit

Transfer Function

E I
1/Z(s)

Z(s) = L1s + L2s + R1 + R2 + 1/C1s + 1/C2s

Transfer Function for Parallel Circuit

General Series Circuit

author w chiyabu Page 20


Transfer Function
E I
1/Z(s)

1
Z(s) =
1 / L1s + 1 / L2s + 1 / R1 + 1 / R2 + C1s + C2s

Example: Let it be desired to determine the equation relating the output voltage E2
to the input voltage E1 in the circuit below.

Solution:

Z1(s) = 1 / [1/R1 + C1s] Z2(s) = R2

Total impedance Z(s) = Z1(s) + Z2(s) = (1 / [1/R1 + C1s]) + R2

Transfer Function

E1 R2(1 + R1C1s) E2
R1 + R2 + R1R2C1s

It can be shown that E2/E1 = R2(1 + R1C1s) / [R1 + R2 + R1R2C1s]

Thus the Transfer Function G(s) = R2(1 + R1C1s) / [R1 + R2 + R1R2C1s]

author w chiyabu Page 21


Operational Amplifiers

An operational amplifier is an electronic amplifier used as a basic building block to


implement transfer functions.

operational amplifier inverting operational amplifier

Characteristics of Operational Amplifiers


1. Differential input, v2(t) - v1(t)
2. High input impedance, Zi = ∞ (ideal)
3. Low output impedance, Zo = 0 (ideal)
4. High constant gain amplification, A = ∞ (ideal)

The output, vo(t), is given by vo(t) = Av2(t) - v1(t)

Inverting Operational Amplifier


If v2(t) is grounded, the amplifier is called an inverting operational amplifier.

For the inverting operational amplifier, we have vo(t) = -Av1(t)

Inverting Operational Amplifier for Transfer Function Realisation

If the input impedance to the amplifier is high, then by Kirchhoff’s current law

Ia(s) = 0 and I1(s) = -I2(s)

Also, since the gain A is large, v1(t) ≈ 0.


Thus, I1(s) = Vi(s) / Z1(s), and -I2(s) = -Vo(s) / Z2(s).

author w chiyabu Page 22


Equating the two currents, Vo(s) / Z2(s) = -Vi(s) / Z1(s), or
the Transfer Function of the inverting operational amplifier configured as shown
above is

Example: Find the transfer function, Vo(s) / Vi(s), for the circuit given below

Solution

After simplifying Z2(s) / Z1(s) we have

Translational Mechanical System Transfer Functions

Mechanical systems parallel electrical networks to such an extent that there are
analogies between electrical and mechanical components and variables. Mechanical
systems, like electrical networks, have three passive, linear components:
The spring and the mass are energy-storage elements. The viscous damper
dissipates energy. The two energy-storage elements are analogous to the two
electrical energy-storage elements, the inductor and capacitor. The energy dissipater
is analogous to electrical resistance.

author w chiyabu Page 23


Force-velocity, force-displacement, and impedance translational relationships
for springs, viscous dampers, and mass

K = spring constant, fv = coefficient of viscous friction, and M = mass

The mechanical system requires just one differential equation, called the equation of
motion, to describe it.

Procedure for Determining Mechanical System Transfer Function


Assume a positive direction of motion to the right. Using our assumed direction of
positive motion,

1. We draw a free-body diagram, placing on the body all forces that act on the body
either in the direction of motion or opposite to it.
2. We use Newton’s law to form a differential equation of motion by summing the
forces and setting the sum equal to zero.
3. Assuming zero initial conditions, we take the Laplace transform of the differential
equation, separate the variables, and arrive at the transfer function.

Example: Find the transfer function, X(s)/F(s) for the system below

author w chiyabu Page 24


Solution
1. Draw the free-body diagram

Free-body diagram of mass, spring, transformed free-body diagram


and damper system

Place on the mass all forces felt by the mass. We assume the mass is travelling
toward the right. Thus, only the applied force points to the right; all other forces
impede the motion and act to oppose it. Hence, the spring, viscous damper, and the
force due to acceleration point to the left.

2. Now write the differential equation of motion using Newton’s law to sum to zero
all of the forces shown on the mass

3. Taking the Laplace transform, assuming zero initial conditions

Ms2X(s) + fvsX(s) + KX(s) = F(s) or

. (Ms2 + fvs + K)X(s) = F(s)

Solving for the transfer function yields

block diagram

General form of equation: [Sum of impedances]X(s) = [Sum of applied forces]

In mechanical systems, the number of equations of motion required is equal to the


number of linearly independent motions. Linear independence implies that a
point of motion in a system can still move if all other points of motion are held still.
Another name for the number of linearly independent motions is the number of
degrees of freedom.
In a mechanical system with two degrees of freedom, one point of motion can be
held still while the other point of motion moves under the influence of an applied
force.

In order to work such a problem, we draw the free-body diagram for each point of
motion and then use superposition. For each free-body diagram we begin by holding
all other points of motion still and finding the forces acting on the body due only to
its own motion. Then we hold the body still and activate the other points of motion
one at a time, placing on the original body the forces created by the adjacent
motion.

author w chiyabu Page 25


Using Newton’s law, we sum the forces on each body and set the sum to zero. The
result is a system of simultaneous equations of motion. As Laplace transforms, these
equations are then solved for the output variable of interest in terms of the input
variable from which the transfer function is evaluated.

Example: Find the transfer function, X2(s) =F(s), for the system below

Two degrees-of-freedom translational mechanical system


Solution

Considering M1 only

Forces on M1 forces on M1 all forces on M1


due only to motion of M1 due only to motion of M2

Considering M2 only

Forces on M2 due only forces on M2 due only all forces on M2


to motion of M2 to motion of M1

The Laplace transform of the equations of motion can now be written

[M1s2(Fv1 + fv3)s + (K1 + K2)]X1(s) – (fv3 s + K2)X2(s) = F(s)


- (fv3 s + K2)X1(s) + [M2s2 + (fv2 + fv3)s + (K2 + K3)]X2(s) = 0

author w chiyabu Page 26


From this, the transfer function, X2(s)/F(s), is

where

block diagram

Equations of Motion by Inspection

Writing equations of motion of a three-degrees-of-freedom mechanical network


by inspection, without drawing the free-body diagram

Example: Write, but do not solve, the equations of motion for the mechanical
network below

The system has three degrees of freedom, since each of the three masses can be
moved independently while the others are held still.

For M1

author w chiyabu Page 27


For M2

For M3

M1 has two springs, two viscous dampers, and mass associated with its motion.
There is one spring between M1 and M2 and one viscous damper between M1 and
M3.

For M1: [M1s2 + (fv1 + fv3)s + (K1 + K2)]X1(s) - K2X2(s) - fv3sX3(s) = 0

For M2 : - K2X1(s) + [M2s2 + (fv2 + fv4)s + K2]X2(s) - fv4sX3(s) = F(s)

For M3: -fv3sX1(s) - fv4sX2(s) + [M3s2 + (fv3 + fv4)s]X3(s) = 0

We can solve these equations for any displacement, X1(s); X2(s); or X3(s), or transfer
function.

author w chiyabu Page 28


Mechanical Components

Series Mechanical Elements

mechanical elements in series

In determining inertia force the acceleration of a mass is always taken with respect
to the earth.
For series mechanical elements, the force f is equal to the summation of the forces
acting on each individual component, and each element undergoes the same
displacement.

Thus f = (K1 + K2 + C1s + C2s + Ms2)x

where x and f are measured from a convenient reference point.

The equivalent impedance for mechanical elements in series is

Z(s) = K1 + K2 + C1s + C2s + Ms2

Transfer Function
f x
1/Z(s)

Parallel Mechanical Elements

mechanical elements in parallel

author w chiyabu Page 29


For parallel mechanical elements, the force f is transmitted through each element. In
addition the deflection x is seen to be the sum of the individual deflections of each
element.

Thus x = (1 / K1 + 1 / K2 + 1 / C1s + 1 / C2s) * f

or f= 1 * x = Zx
1 / K1 + 1 / K2 + 1 / C1s + 1 / C2s

The equivalent impedance for mechanical elements in parallel is

Z(s) = 1
1 / K1 + 1 / K2 + 1 / C1s + 1 / C2s

Transfer Function
f x
1/Z(s)

A necessary condition for parallel elements is that the force be transmitted through
each element. Springs and dampers satisfy this condition because the force is the
same on both sides. However, this is not the case for a mass such as the one in (a)
in the figure below, because the difference in forces acting on both sides of a mass
is utilised in acceleration. Thus, a mass located between other elements cannot be in
parallel with them.

mechanical system

author w chiyabu Page 30


A mass can be in parallel only if it is the last element as shown below.

parallel mass-spring-damper combination

For this system, the displacement x is

x = (x – y) + (y – z) + z = (1/K + 1/Cs + 1/Ms2) * f

Rotational Mechanical System Transfer Functions

Rotational mechanical systems are handled the same way as translational


mechanical systems, except that torque replaces force and angular displacement
replaces translational displacement. The mechanical components for rotational
systems are the same as those for translational systems, except that the
components undergo rotation instead of translation.

Note: K = spring constant, D = coefficient of viscous friction, and J = moment of inertia

Writing the equations of motion for rotational systems is similar to writing them for
translational systems; the only difference is that the free-body diagram consists of
torques rather than forces.
We obtain the torques using superposition.
1. We rotate a body while holding all other points still and place on its free-body
diagram all torques due to the body’s own motion.
2. Then, holding the body still, we rotate adjacent points of motion one at a time
and add the torques due to the adjacent motion to the free-body diagram.
3. The process is repeated for each point of motion.
4. For each free-body diagram, these torques are summed and set equal to zero to
form the equations of motion.

author w chiyabu Page 31


Example:
Find the transfer function, θ2(s)/T(s), for the rotational system shown below. The
rod is supported by bearings at either end and is undergoing torsion. A torque is
applied at the left, and the displacement is measured at the right.

Solution:
1. Obtain the schematic from the physical system. Even though torsion occurs
throughout the rod, we approximate the system by assuming that the torsion acts
like a spring concentrated at one particular point in the rod, with an inertia J1 to
the left and an inertia J2 to the right. We also assume that the damping inside the
flexible shaft is negligible.

There are two degrees of freedom, since each inertia can be rotated while the
other is held still. Hence, it will take two simultaneous equations to solve the
system.

2. Draw a free-body diagram of J1, using superposition.

With J2 held still

(a) shows the torques on J1 if J2 is held still and J1 rotated. (b) shows the torques
on J1 if J1 is held still and J2 rotated. Finally, (c) is the final free-body diagram for
J1. The same process is repeated in Figure 2.24 for J2.

With J1 held still

author w chiyabu Page 32


Summing torques, respectively we obtain the equations of motion,

(J1s2 + D1s + K.)θ1(s) - Kθ2(s) = T(s)

-Kθ1(s) + (J2s2 + D2s + K)θ2(s) = 0

Required transfer function is θ2(s)/T(s) = K / ∆

where

General Form:

Block Diagram

Analogy between Electrical and Mechanical Control Systems

An electric circuit that is analogous to a system from another discipline is called an


electric circuit analogue. Analogues can be obtained by comparing the describing
equations, such as the equations of motion of a mechanical system, with either
electrical mesh or nodal equations.
When compared with mesh equations, the resulting electrical circuit is called a series
analogue. When compared with nodal equations, the resulting electrical circuit is
called a parallel analogue.

author w chiyabu Page 33


Series Analogue

Example:

mechanical system desired electrical representation series analogue

Parameters for Series Analogue

mass = M inductor = M henries


viscous damper = fv resistor = fv ohms
spring = K capacitor = 1/K farads
applied force = f(t) voltage source = f(t)
velocity = v(t) mesh current = v(t)

Translational Mechanical System Equation:

(Ms2 + fvs + K)X(s) = F(s)

From Kirchhoff’s Laws for the series RLC network

(Ls + R + 1/Cs)I(s) = E(s)

Electrical analogue

Example: Draw a series analogue for the mechanical system below

author w chiyabu Page 34


Solution:
Electrical analogue

Parallel Analogue
A system can also be converted to an equivalent parallel analogue.

Example:

mechanical system desired electrical representation parallel analogue

Parameters for Parallel Analogue

mass = M capacitor = M farads


viscous damper = fv resistor = 1/fv ohms
spring = K inductor = 1/K henries
applied force = f(t) current source = f(t)
velocity = v(t) node voltage = v(t)

From Kirchhoff’s Laws for the series RLC network

(Cs + 1/R + 1/Ls)E(s) = I(s)

author w chiyabu Page 35


Analogous Quantities in a Direct (Force-Voltage) Analogue

translational force mass viscous damping spring displacement velocity


mechanical coefficient constant
system f M B K x á= sx
electrical voltage inductance resistance reciprocal of charge current
system capacitance
E L R 1/C Q I = sQ

Analogous Quantities in an Inverse (Force-Current) Analogue

translational force velocity spring damping mass


mechanical constant coefficient
system f á K B M
electrical current voltage reciprocal of reciprocal of capacitance
system inductance resistance
I E 1/L 1/R C

4. Block Diagram Algebra and Transfer Functions of Systems

Block Diagrams
A Block Diagram of a system is a pictorial representation of the functions performed
by each component and of the flow of signals. Such a diagram depicts the
interrelationships that exist among the various components. A Block Diagram is
shorthand, graphical representation of a physical system, illustrating the functional
relationships among its components.
Block Diagrams permit evaluation of the contributions of the individual elements to
the overall performance of the system.

In a block diagram all system variables are linked to each other through functional
blocks. The functional block or simply block is a symbol for the mathematical
operation on the input signal to the block that produces the output. The transfer
functions of the components are usually entered in the corresponding blocks, which
are connected by arrows to indicate the direction of the flow of signals.

Canonical Form of a Feedback Control System

author w chiyabu Page 36


The two blocks in the forward path of the feedback system may be combined.
Letting G = G1G2, the resulting configuration is called the canonical form of a
feedback control system.

Definitions
1. G = direct transfer function = forward transfer function
2. H = feedback transfer function
3. GH = loop transfer function = open-loop transfer function
4. C/R = closed-loop transfer function = control ratio
5. E/R = actuating signal ration = error ratio
6. B/R = primary feedback ratio

In the following equations, the minus (-) sign refers to a positive feedback system,
and the plus (+) sign refers to a negative feedback system.

The characteristic equation of the system, which is determined from 1 ± GH = 0,


is
DGH ± NGH = 0

where DGH is the denominator and NGH the numerator of GH

author w chiyabu Page 37


Block Diagram Transformation Theorems
In the tables below the letter P is used to represent any transfer function, and W, X,
Y, Z denote any s-domain signals.

author w chiyabu Page 38


Unity Feedback Control Systems
A unity feedback control system is a feedback system in which the primary
feedback b is identically equal to the controlled output c.

Example: H = 1 for a linear, unity feedback system

R E C
G
±

Any feedback system with only linear elements in the feedback loop can be put into
the form of a unity feedback system.

Example:

The characteristic equation for the unity feedback system, determined from 1±G = 0
is
DG ± NG = 0

where DG is the denominator and NG the numerator of G

author w chiyabu Page 39


Multiple Inputs
Sometimes it is necessary to evaluate a system’s performance when several stimuli
are simultaneously applied at different points of the system.
When multiple inputs are present in a linear system, each is treated independently of
the others. The output due to all stimuli acting together is found in the following
manner.

Step 1: Set all inputs except one equal to zero


Step 2: Transform the block diagram to canonical form
Step 3: Calculate the response due to the chosen input acting alone
Step 4: Repeat steps 1 to 3 for each of the remaining inputs
Step 5: Algebraically add all the responses (outputs) determined in steps 1 to 5. The
sum is the total output of the system with all inputs acting simultaneously.

Example: Determine the output C of the following system.

Step 1: Put U = 0
Step 2: The system reduces to

Step 3: The out CR due to input R is CR = [G1G2 / (1 + G1G2)] * R


Step 4a: Put R = 0
Step 4b: Put -1 into a block, representing the negative feedback effect

Rearrange the block diagram

author w chiyabu Page 40


Let the -1 block be absorbed into the summing point

Step 4c: The output CU due to input U is CU = [G2 / (1 + G1G2)] U


Step 5: The total output is

Reduction of Complicated Block Diagrams


The block diagram of a practical feedback control system is often quite complicated.
It may include several feedback or feedforward loops, and multiple inputs. By means
of systematic block diagram reduction, every multiple loop feedback system may be
reduced to canonical form.
The following general steps may be used as a basic approach in the reduction of
complicated block diagrams. Each step refers to a specific transformation in the
chart given above.

Step 1: Combine all cascade blocks using Transformation 1


Step 2: Combine all parallel blocks using Transformation 2
Step 3: Eliminate all minor feedback loops using Transformation 4
Step 4: Shift summing points to the left and takeoff points to the right of the major
loop, using Transformation 7, 10, and 12
Step 5: Repeat steps 1 to 4 until the canonical form has been achieved for a
particular input
Step 6: Repeat steps 1 to 5 for each input as required

Transformations 3, 5, 6, 8, 9, and 11 are sometimes useful, and experience with the


reduction technique will determine their application.

Example: Reduce the following block diagram to canonical form

author w chiyabu Page 41


Solution:

Step 1:

Step 2:

Step 3:

Step 4: Does not apply

Step 5:

Step 6: Does not apply

An occasional requirement of block diagram reduction is the isolation of a particular


block in a feedback or feedforward loop. This may be desirable in order to more
easily examine the effect of a particular block on the overall system.
Isolation of a block may generally be accomplished by applying the same reduction
steps to the system, but usually in a different order. Also, the block to be isolated
cannot be combined with any others.
Rearranging Summing Points (Transformation 6), and Transformations 8, 9, and 11
are especially useful for isolating blocks.

Example: Reduce the following block diagram, isolating block H1

author w chiyabu Page 42


Solution:

Steps 1 and 2

We do not apply step 3 at this time, but go directly to step 4, moving takeoff point 1
beyond block G2 + G3

We may now rearrange summing points 1 and 2 and combine the cascade blocks in
the forward loop using Transformation 6, then, 1

Step 3:

author w chiyabu Page 43


Finally, we apply Transformation 5 to remove [1 / (G2 + G3)]

Note that the same result could have been obtained after applying step 2 by moving
takeoff point 2 ahead of G2 + G3, instead of takeoff point 1 beyond G2 + G3.
Block G2 + G3 has the same effect on the control ratio C/R whether it directly follows
R or directly precedes C.

5. Signal-Flow Graphs

A Signal-Flow Graph is a pictorial representation of the simultaneous equations


describing a system. It graphically displays the transmission of signals through the
system, as does the block diagram.
Signal-Flow Graphs are an alternative to block diagrams. Unlike block diagrams,
which consist of blocks, signals, summing junctions, and pickoff points, a signal-flow
graph consists only of branches, which represent systems, and nodes, which
represent signals.

Signal-Flow Graph Components

system signal interconnection of systems and signals

A system is represented by a line with an arrow showing the direction of signal


flow through the system. Adjacent to the line we write the transfer function. A
signal is a node with the signal’s name written adjacent to the node. Each signal is
the sum of signals flowing into it.

Example:
signal V(s) = R1(s)G1(s) - R2(s)G2(s) + R3(s)G3(s)
signal C2(s) = V(s)G5(s) = R1(s)G1(s)G5(s) - R2(s)G2(s)G5(s) + R3(s)G3(s)G5(s)
signal C3(s) = -V(s)G6(s) = - R1(s)G1(s)G6(s) + R2(s)G2(s)G6(s) - R3(s)G3(s)G6(s).

Not: in summing negative signals we associate the negative sign with the system
and not with a summing junction

author w chiyabu Page 44


5.1 Signal-Flow Graph Algebra

1. The Addition Rule


The value of a node designated by a node is equal to the sum of all signals
entering the node.

Example: The signal-flow graph for the equation of a line in rectangular


coordinates, Y = mX + b, is
X X
m m
Y or Y
1 b
b 1

2. The Transmission Rule


The value of the variable designated by a node is transmitted on every branch
leaving the node.

Example: The signal-flow graph of the simultaneous equations Y = 3X, Z = - 4X is


3 Y

X
-4 Z

3. Multiplication Rule
A cascade (series) connection of n-1 branches with transmission functions
A21, A32, A43 ... An(n01) can be replaced by a single branch with a new transmission
function equal to the product of the old ones.

Example: The signal-flow graph of simultaneous equations Y = 10X, Z = -20Y is

10 -20 -200
reduces to
X Y Z X Z

4. The value of the variable represented by a node is equal to the sum of all the
signals entering the node.

node as a summing point and as a transmitting point

author w chiyabu Page 45


For the signal-flow graph above, the value of y1 is equal to the sum of the signals
transmitted through all the incoming branches.

That is y1 = a21y2 + a31y3 + a41y4 + a51y5

5. The value of the variable represented by a node is transmitted through all the
branches leaving the node.
For the signal-flow graph above, y6 = a16y1 y7 = a17y1 y8 = a18y1

6. Parallel branches in the same direction connected between two nodes can be
replaced by a single branch with gain equal to the sum of the gains of the parallel
branches.

Example:

7. A series connection of unidirectional branches can be replaced by a single branch


with gain equal to the product of the branch gains.

Example:

8. Signal-flow graph of a feedback control system

author w chiyabu Page 46


The signal-flow graph may be regarded as a simplified notation for the block
diagram.
Writing the equations for the nodes at E(s) and C(s) we have

E(s) = R(s) – H(s)C(s) and C(s) = G(s)H(s)

Substituting for E(s) we have the closed-loop transfer function as

Examples of signal-flow graph simplifications

parallel branches series branches

star branches reduction of feedback loop to self-loop

reduction of self-loop: conventional graph reduction of self-loop: regular algebra

removal of a node with self-loop a node considered as a source and sink

author w chiyabu Page 47


In conventional flow-graph theory, a self-loop corresponds to division.
For |g| < 1,

In regular expression algebra, this is analogous to g* = λ + g + g2 + g3 +...

In simplifying a graph with self-loop, the loop can be replaced by a branch with gain
b*, and the node can be removed by adjusting the gain of all the paths through the
node by b*.

5.2 Definitions

The following terminology is associated with the following signal-flow graph.

1. A path is a continuous, unidirectional succession of branches along which no


node is passed more than once.
Example: X1 to X2 to X3 to X4; X2 to X3 and back to X2.

2. An input node or source is a node with only outgoing branches.


Example: X1 is an input node

3. An output node or sink is a node with only incoming branches.


Example: X4 is an output node

4. A forward path is a path from the input node to the output node.
Example: X1 to X2 to X3 to X4 and X1 to X2 to X4 are forward paths.

5. A feedback path or feedback loop is a path which originates and terminates


on the same node.
Example: X2 to X3 and back to X2 is a feedback path.

6. A self-loop is a feedback loop consisting of a single branch.


Example: A33 is a self-loop.

7. The gain of a branch is the transmission function of that branch when the
transmission function is a multiplicative operator.
Example: A33 is the gain of the self-loop if A33 is a constant or transfer function.

author w chiyabu Page 48


8. The path gain is the product of the branch gains encountered in traversing a
path.
Example: The path gain of the forward path
from X1 to X2 to X3 to X4 is A21A32A43

9. The loop path is the product of the branch gains of the loop.
Example: The loop gain of the feedback loop from X2 to X3 and back to X2 is
A32A23.

In the canonical feedback system, if the signal-flow graph were to be drawn directly
from the equations, the “output node” would require an outgoing branch, contrary
to definition. This situation may be remedied by adding a branch with a transmission
function of unity entering a “dummy” node.

Example

5.3 Construction of Signal-Flow Graphs

The signal-flow graph of a linear feedback control system whose components are
specified by non-interacting transfer functions can be constructed by direct reference
to the block diagram of the system. Each variable of the block becomes a node
and each block becomes a branch.

Example:
The block diagram of the canonical feedback control system is given by

The signal-flow diagram is constructed from the block diagram

Note that the – or + sign of the summing point is associated with H.

author w chiyabu Page 49


General Procedure
The signal-flow graph of a system described by a set of simultaneous equations can
be constructed in the following manner.

1. Write the system equation in the form


X1 = A11X1 + A12X2 + ... + A1nXn
X2 = A21X1 + A22X2 + ... + A2nXn
.............................................
Xm = Am1X1 + Am2X2 + ... + AmnXn

An equation for X1 is not required if X1 is an input node.

2. Arrange the m or n (whichever is larger) nodes from left to right. The nodes may
be arranged if the required loops later appear too cumbersome.

3. Connect the nodes by the appropriate branches A11, A12, etc

4. If the desired output node has outgoing branches, add a dummy node and a
unity-gain branch.

5. Rearrange the nodes and/or loops in the graph to achieve maximum pictorial
clarity.

Example: Construct a signal-flow graph for the resistance network below.

Solution:
There are five variables: v1, v2, v3, i1 and i2. Using Kirchhoff’s Voltage and Current
Laws, we can write four independent equations. Proceeding from left to right in the
schematic, we have

Laying out the five nodes in the same order with v1 as an input node, and
connecting the nodes with the appropriate branches, we have

author w chiyabu Page 50


If we wish to consider v3 as an output node, we add a unity-gain branch and
another node, yielding

Properties of Signal-Flow Graphs

1. A signal-flow graph applies only to linear systems.


2. The equations from which a signal-flow graph is drawn must be algebraic
equations in the form of effects in form of causes.
3. Nodes are used to represent variables. Normally, the nodes are arranged from left
to right, following a succession of causes and effects through the system.
4. Signals travel along branches only in the direction described by the arrows of the
branches.
5. The branch directing from node yk to yj represents the dependence of the variable
yj upon yk, but not the reverse.
6. A signal yk travelling along a branch between nodes yk and yj is multiplied by the
gain of the branch, akj, so that signal akjyk is delivered at node yj.

5.4 The General Input-Output Gain Formula

We can reduce complicated block diagrams to canonical form and obtain the the
following control ratio

It is possible to simplify signal-flow graphs in a similar manner to that of block


diagrams. It is also possible, and much less time-consuming, to write down the
input-output relationship by inspection from the original signal-flow graph.

Let us denote the ratio of the input variable to the output variable by T. For linear
feedback control systems, T = C/R.

author w chiyabu Page 51


The general formula for any signal-flow graph is

where Pi = the ith forward path gain


Pjk = jth possible product of k nontouching loop gains

∆ = 1 – (sum of loop gains) + (sum of all gain products of two nontouching loops)
- (sum of all gain products of three nontouching loops) + ...
∆1 = ∆ evaluated with all loops touching Pi eliminated

Two loops, paths, or a loop and a path are said to be nontouching if they have no
nodes in common.
∆ is called the signal-flow graph determinant or characteristic function, since
∆ = 0 is the system characteristic equation.

Example:
Determine C/R for the signal-flow graph of the following canonical feedback system

Solution:
There is only one forward path. Hence

P1 = G P2 = P3 = ... = 0

There is only one forward feedback path. Hence

P11 = GH Pjk = 0, j ≠ 1, k ≠ 1
Then ∆ = 1 – P11 = 1 ± GH and ∆1 = 1 – 0 = 1

Finally,

author w chiyabu Page 52


Transfer Function Computation of CASCADED Components
Loading effects of interacting components require little special attention using signal-
flow graphs. Simply combine the graphs of the components at their normal joining
points (input node of one to the input node of another), account for loading by
adding new loops at the joined nodes, and compute the overall gain.

Example:
Assume that two identical resistance networks are to be cascaded and used as the
control elements in the forward loop of a control system. The networks are voltage
dividers of the form

The independent equations for this network are

Signal-Flow Graph:

The gain of this network, by inspection, equal v2/v1 = R3 /(R1 + R3)

For Cascaded Network:

There are five variables: v1, v2, v3, i1 and i2. Using Kirchhoff’s Voltage and Current
Laws, we can write four independent equations. Proceeding from left to right in the
schematic, we have

author w chiyabu Page 53


Laying out the five nodes in the same order with v1 as an input node, and
connecting the nodes with the appropriate branches, we have

If we wish to consider v3 as an output node, we add a unity-gain branch and


another node, yielding

We observe that the feedback branch –R3 in the above graph does not appear in the
signal-flow graph of the cascaded signal-flow graphs of the individual networks
connected from v2 to v’1:

This means that as a result of connecting the two networks, the second one loads
the first, changing the equation for v2 from

v2 = R3i1 to v2 = R3i1 – R3i2

The gain of the combined network is

It is good general practice to calculate the gain of cascaded networks directly from
the combined signal-flow graph. Most practical control system components load each
other when connected in series.

author w chiyabu Page 54


5.5 Block Diagram Reduction using Signal-Flow Graphs and
The General Input-Output Gain Formula

Often, the easiest way to determine the control ratio of a complicated block diagram
is to translate the block diagram into a signal-flow graph. Takeoff points and
summing points must be separated by a unity-gain branch in the signal-flow graph.
If the elements of G and H of a canonical representation are desire, the direct
transfer function is

The loop transfer function is GH = ∆ - 1


The two equations are solved simultaneously for G and H, and the cononical
feedback control system is drawn from the result.

Example: Determine the control ratio C/R and the canonical diagram of the feedback
control system below

The signal-flow graph is

There are two forward paths: P1 = G1G2G4, P1 = G1G3G4


There are three feedback loops: P11 = G1G4H1, P21 = -G1G2G4H2, P31 = -G1G3G4H2
There no nun-touching loops, and all loops touch both forward paths; then
∆1 = 1, ∆2 = 1
Therefore, the control ratio is:

author w chiyabu Page 55


Now G = G1G4(G2 + G3) and GH = G1G4(G3H2 + G2H2 – H1)
Therefore,

The canonical block diagram is given by

5.6 Converting Common Block Diagrams to Signal-Flow Graphs

Example: Convert the cascaded, parallel, and feedback forms of the block diagrams
shown in (a), (b), and (c), respectively, into signal-flow graphs.

(a)

(b)

(c)

author w chiyabu Page 56


Solution:
In each case, we start by drawing the signal nodes for that system. Next we
interconnect the signal nodes with system branches.

(a)

cascaded system nodes cascaded system signal-flow graph

(b)

parallel system nodes parallel system signal-flow graph

(c)

author w chiyabu Page 57


feedback system nodes feedback system signal-flow graph

Example: Convert the block diagram below to a signal-flow graph

Solution:
Begin by drawing the signal nodes. Next, interconnect the nodes, showing the
direction of signal flow and identifying each transfer function.

signal nodes

signal-flow graph

author w chiyabu Page 58


Finally, if desired, simplify the signal-flow graph to the one shown below by
eliminating signals that have a single flow in and a single flow out, such as V2(s),
V6(s), V7(s), and V8(s).

simplified signal-flow graph

5.7 Signal-Flow Graphs of State Equations

Two approaches are available for the analysis and design of feedback control
systems:

1. The Classical, or frequency-domain, Technique


This approach is based on converting a system’s differential equation to a transfer
function, thus generating a mathematical model of the system that algebraically
relates a representation of the output to a representation of the input. Replacing
a differential equation with an algebraic equation not only simplifies the
representation of individual subsystems but also simplifies modelling
interconnected subsystems.
The primary disadvantage of the classical approach is its limited applicability:
It can be applied only to linear, time-invariant systems or systems that can be
approximated as such.
A major advantage of frequency-domain techniques is that they rapidly provide
stability and transient response information. Thus, we can immediately see the
effects of varying system parameters until an acceptable design is met.

2. The State-Space Approach (also referred to as the Modern, or time-domain,


approach) is a unified method for modelling, analysing, and designing a wide
range of systems.
For example, the state-space approach can be used to represent nonlinear
systems that have backlash, saturation, and dead zone. Also, it can handle,
conveniently, systems with nonzero initial conditions.
The time-domain approach can be used to represent systems with a digital
computer in the loop or to model systems for digital simulation. With a simulated
system, system response can be obtained for changes in system parameters—an
important design tool. The state-space approach is also attractive because of the
availability of numerous state-space software packages for the personal
computer.
The time-domain approach can also be used for the same class of systems
modelled by the classical approach. This alternate model gives the control
systems designer another perspective from which to create a design.

author w chiyabu Page 59


While the state-space approach can be applied to a wide range of systems, it is not
as intuitive as the classical approach. The designer has to engage in several
calculations before the physical interpretation of the model is apparent, whereas in
classical control a few quick calculations or a graphic presentation of data rapidly
yields the physical interpretation.

6. Mason’s Rule

Mason’s Rule for reducing a signal-flow graph to a single transfer function requires
the application of one formula. The formula was derived by S. J. Mason when he
related the signal-flow graph to the simultaneous equations that can be written from
the graph (Mason, 1953).

Gives overall transfer function in terms of:


1. Forward path gains (no vertices repeat).
2. Loop gains (path starts and ends at same node).
3. Products of loop gains for nontouching loops.
4. Nontouching loops: no common nodes.
5. Two nontouching loops, three nontouching loops, ….

6.1 Definitions

Loop Gain: The product of branch gains found by traversing a path that starts at a
node and ends at the same node, following the direction of the signal flow, without
passing through any other node more than once.

Example

signal-flow graph for demonstrating Mason’s rule

There are four loop gains:


1. G2(s)H1(s) 2. G4(s).H2(s) 3. G4(s)G5(s)H3(s) 4. G4(s)G6(s)H3(s)

Forward-path Gain: The product of gains found by traversing a path from the
input node to the output node of the signal-flow graph in the direction of signal flow.
There are two forward-path gains:
1. G1(s)G2(s)G3(s)G4(s)G5(s)(s)G7(s)
2. G1(s)G2(s)G3(s)G4(s)G6(s)G7(s)

author w chiyabu Page 60


Nontouching Loops: Loops that do not have any nodes in common.
Loop G2(s)H1(s) does not touch loops G4(s)H2(s), G4(s)G5(s)H3(s), and
G4(s)G6(s)H3(s)

Nontouching-loop Gain: The product of loop gains from nontouching loops taken
two, three, four, or more at a time.

The product of loop gain G2(s)H1(s) and loop gain G4(s)H2(s) is a nontouching-loop
gain taken two at a time.
In summary, all three of the nontouching-loop gains taken two at a time are

1. [G2(s)H1(s)] [G4(s)H2(s)]
2. [G2(s)H1(s)] [G4(s)G5(s)H3(s)]
3. [G2(s)H1(s)] [G4(s)G6(s)H3(s)]

The product of loop gains [G4(s)G5(s)H3(s)] [G4(s)G6(s)H3(s)] is not a nontouching-


loop gain since these two loops have nodes in common.
In this example there are no nontouchingloop gains taken three at a time since three
nontouching loops do not exist in the example.

6.2 Mason’s Rule


The transfer function, C(s)/R(s), of a system represented by a signal-flow graph is

eq1
where
k = number of forward paths
Tk = the kth forward-path gain

∆ = 1 - Σ(loop gains) + Σ(nontouching-loop gains taken two at a time)


– Σ(nontouching-loop gains taken three at a time)
+ Σ(nontouching-loop gains taken four at a time) - . . .

∆k = ∆ – Σ(loop gain terms in ∆ that touch the kth forward path).


In other words, ∆k is formed by eliminating from ∆ those loop gains that touch the
kth forward path.

Notice the alternating signs for the components of ∆.

author w chiyabu Page 61


6.3 Transfer Function via Mason’s Rule

Example: Find the transfer function, C(s)/R(s), for the signal-flow graph below.

Solution

1. Identify the forward-path gains.


In this example there is only one: G1(s)G2(s)G3(s)G4(s)G5(s) eq2

2. Identify the loop gains.


There are four, as follows:
1. G2(s)H1(s) 2. G4(s)H2(s) 3. G7(s)H4(s) 4. G2(s)G3(s)G4(s)G5(s)G6(s)G7(s)G8(s)

3. Identify the nontouching loops taken two at a time.


Loop 1 does not touch Loop 2, loop 1 does not touch Loop 3, and Loop 2 does
not touch Loop 3. Loops 1, 2, and 3 all touch Loop 4.

Thus, the combinations of nontouching loops taken two at a time are as follows:
Loop 1 and loop 2: G2(s)H1(s)G4(s)H2(s)
Loop 1 and loop 3: G2(s)H1(s)G7(s)H4(s)
Loop 2 and loop 3: G4(s)H2(s)G7(s)H4(s)
4. The nontouching loops taken three at a time are as follows:
Loops 1, 2, and 3: G2(s)H1(s)G4(s)H2(s)G7(s)H4(s)

Now, from Mason’s Rule, we form ∆ and ∆k.


Hence,
∆ = 1- [G2(s)(s)H1(s) + G4(s)H2(s) + G7(s)H4(s)
+ G2(s)G3(s)G4(s)G5(s)G6(s)G7(s)G8(s)]
+ [G2(s)H1(s)G4(s)H2(s) + G2(s)H1(s)G7(s)H4(s) + G4(s)H2(s)G7(s)H4(s)]
- [G2(s)H1(s)G4(s)H2(s)G7(s)H4(s)] eq3

We form ∆k by eliminating from ∆ the loop gains that touch the kth forward path:
∆1 = 1 - G7(s)H4(s) eq4

author w chiyabu Page 62


Expressions (eq2), (eq3), and (eq4) are now substituted into (eq1), yielding the
transfer function:

Since there is only one forward path, G(s) consists of only one term, rather than a
sum of terms, each coming from a forward path.

7.0 Time Response of Control Systems:


Poles, Zeros, and System Response

7.1 Introduction
The output response of a system is the sum of two responses: the forced
response and the natural response. Although many techniques, such as solving a
differential equation or taking the inverse Laplace transform, enable us to evaluate
this output response, these techniques are laborious and time-consuming.
Productivity is aided by analysis and design techniques that yield results in a
minimum of time. If the technique is so rapid that we feel we derive the desired
result by inspection, we sometimes use the attribute qualitative to describe the
method. The use of poles and zeros and their relationship to the time response of a
system is such a technique.
The concept of poles and zeros, fundamental to the analysis and design of control
systems, simplifies the evaluation of a system’s response.

Poles of a Transfer Function


The poles of a transfer function are:
(1) the values of the Laplace Transform variable, s, that cause the Transfer Function
to become infinite or
(2) any roots of the denominator of the Transfer Function that are common to roots
of the numerator.

Generally, poles are defined as the roots of the denominator of a transfer function.

Strictly speaking, the poles of a transfer function satisfy part (1) of the definition.
For example, the roots of the characteristic polynomial in the denominator are
values of s that make the transfer function infinite, so they are thus poles. However,
if a factor of the denominator can be cancelled by the same factor in the numerator,
the root of this factor no longer causes the transfer function to become infinite. In
control systems, we often refer to the root of the cancelled factor in the
denominator as a pole even though the transfer function will not be infinite at this
value.

author w chiyabu Page 63


Zeros of a Transfer Function
The zeros of a transfer function are:
(1) the values of the Laplace Transform variable, s, that cause the transfer function
to become zero, or
(2) any roots of the numerator of the transfer function that are common to roots of
the denominator.

Generally, zeros are defined as the roots of the polynomial of the numerator of a
Transfer Function.

Strictly speaking, the zeros of a transfer function satisfy part (1) of this definition.
For example, the roots of the numerator are values of s that make the transfer
function zero and are thus zeros. However, if a factor of the numerator can be
cancelled by the same factor in the denominator, the root of this factor no longer
causes the transfer function to become zero. In control systems, we often refer to
the root of the cancelled factor in the numerator as a zero even though the transfer
function will not be zero at this value.

Poles and Zeros of a First-Order System: An Example

system showing input and output pole-zero plot of the system


× = pole = zero

evolution of a system response

author w chiyabu Page 64


From the given function, a pole exists at s = -5, and a zero exists at -2. These
values are plotted on the complex s-plane.

To show the properties of the poles and zeros, let us find the unit step response of
the system. Multiplying the transfer function G(s) by a step function yields

where

Thus,

Characteristics of Poles and Zeros

1. A pole of the input function generates the form of the forced response (that
is, the pole at the origin generated a step function at the output).
2. A pole of the transfer function generates the form of the natural response
(that is, the pole at -5 generated e-5t).
3. A pole on the real axis generates an exponential response of the form e-αt,
where -α is the pole location on the real axis. Thus, the farther to the left a pole is
on the negative real axis, the faster the exponential transient response will decay
to zero (again, the pole at -5 generated e-5t.
4. The zeros and poles generate the amplitudes for both the forced and natural
responses.

effect of a real-axis pole upon transient response

Example: Given the system below, write the output, c(t), in general terms. Specify
the forced and natural parts of the solution.

author w chiyabu Page 65


SOLUTION:
By inspection, each system pole generates an exponential as part of the natural
response. The input’s pole generates the forced response.

Thus,

Taking the inverse Laplace transform, we get

Exercise:
A system has a transfer function

Write, by inspection, the output, c(t), in general terms if the input is a unit step.

Solution: c(t) ≡ A + Be-t + Ce-7t + De-8t + Ee-10t

Summary

Poles determine the nature of the time response:


1. Poles of the input function determine the form of the forced response,
2. Poles of the transfer function determine the form of the natural response.
3. Zeros and poles of the input or transfer function contribute to the amplitudes of
the component parts of the total response.
4. Poles on the real axis generate exponential responses.

author w chiyabu Page 66


7.2 First-Order Systems

We now discuss first-order systems without zeros to define a performance


specification for such a system.

first-order system pole plot

A first-order system without zeros can be described by the transfer function shown
above. If the input is a unit step, where R(s) = 1/s, the Laplace transform of the
step response is C(s), where
C(s) = R(s)G(s) = a / [s(s + a)]

Taking the inverse transform, the step response is given by

c(t) = cf(t) + cn(t) = 1 – e-at

where the input pole at the origin generated the forced response cf(t) = 1, and the
system pole at -a, as shown above, generated the natural response cn(t) = -e-at .

first-order system response to a unit step

author w chiyabu Page 67


Let us examine the significance of parameter a, the only parameter needed to
describe the transient response.

When t = 1/a

or

Time Constant
We call 1/a the time constant of the response. The time constant can be
described as the time for e-at to decay to 37% of its initial value. Alternately, the
time constant is the time it takes for the step response to rise to 63% of its final
value.

The reciprocal of the time constant has the units (1/seconds), or frequency. Thus,
we can call the parameter a the exponential frequency.
Since the derivative of e-at is -a when t = 0, a is the initial rate of change of the
exponential at t = 0.
Thus, the time constant can be considered a transient response specification
for a first-order system, since it is related to the speed at which the system
responds to a step input.
The time constant can also be evaluated from the pole plot. Since the pole of the
transfer function is at -a, we can say the pole is located at the reciprocal of the time
constant, and the farther the pole from the imaginary axis, the faster the transient
response.

Rise Time Tr
Rise time is defined as the time for the waveform to go from 0.1 to 0.9 of its final
value.
Rise time is found by solving c(t) = cf(t) + cn(t) = 1 – e-at for the difference in
time at c(t) = 0.9 and c(t) = 0.1.

Hence,

Settling Time Ts
Settling time is defined as the time for the response to reach, and stay within, 2%
of its final value.
Letting c(t) = 0.98 and solving for time, t, in c(t) = cf(t) + cn(t) = 1 – e-at
we find the settling time to be
Ts = 4 / a

author w chiyabu Page 68


First-Order Transfer Functions via Testing
Often it is not possible or practical to obtain a system’s transfer function analytically.
Perhaps the system is closed, and the component parts are not easily identifiable.
Since the transfer function is a representation of the system from input to output,
the system’s step response can lead to a representation even though the inner
construction is not known. With a step input, we can measure the time constant and
the steady-state value, from which the transfer function can be calculated.

Consider a simple first-order system, G(s) = K /(s + a), whose step response is

If we can identify K and a from laboratory testing, we can obtain the transfer
function of the system.

Example: Assume the unit step response given by c(t) = cf(t) + cn(t) = 1 – e-at.

From the response, we measure the time constant, that is, the time for the
amplitude to reach 63% of its final value. Since the final value is about 0.72, the
time constant is evaluated where the curve reaches 0.63 x 0.72 = 0.45, or about
0.13 second.
Hence, a = 1/0.13 = 7.7.

To find K, we realise from

that the forced response reaches a steady-state value of K/a = 0.72.


Substituting the value of a, we find K = 5.54.
Thus, the transfer function for the system is G(s) = 5.54/(s + 7.7).

The transfer function G(s) = 5/(s + 7) was used to generate the following response.

author w chiyabu Page 69


Exercise:
A system has transfer function G(s) = 50 / (s + 50). Find the time constant, Tc,
settling time, Ts, and rise time, Tr.

Solution: Tc = 0.02 s Ts = 0.08 s Tr = 0.044 s

7.3 Second-Order Systems: Introduction

Compared to the simplicity of a first-order system, a second-order system exhibits a


wide range of responses that must be analyzed and described. Whereas varying a
first-order system’s parameter simply changes the speed of the response, changes in
the parameters of a second-order system can change the form of the response.
For example, a second-order system can display characteristics much like a first-
order system, or, depending on component values, display damped or pure
oscillations for its transient response.

second-order systems, pole plots, and step responses

author w chiyabu Page 70


In the figure above (a) is the general case, which has two finite poles and no zeros.
The term in the numerator is simply a scale or input multiplying factor that can take
on any value without affecting the form of the derived results. By assigning
appropriate values to parameters a and b, we can show all possible second-order
transient responses.
The unit step response then can be found using C(s) = R(s)G(s),
where R(s) = 1/s, followed by a partial-fraction expansion and the inverse Laplace
Transform.

Overdamped Response

overdamped system pole-zero plot response

For this response,

This function has a pole at the origin that comes from the unit step input and two
real poles that come from the system. The input pole at the origin generates the
constant forced response; each of the two system poles on the real axis generates
an exponential natural response whose exponential frequency is equal to the pole
location.
Hence, the output initially could have been written as

c(t) = K1 + K2e-7.854t + K3e-1.146t

This response is called overdamped. We see that the poles tell us the form of the
response without the tedious calculation of the inverse Laplace transform.

Underdamped Response

underdamped system pole-zero plot response

author w chiyabu Page 71


For this response

This function has a pole at the origin that comes from the unit step input and two
complex poles that come from the system. We now compare the response of the
second-order system to the poles that generated it.

1. We will compare the pole location to the time function, and then we will compare
the pole location to the plot.
The poles that generate the natural response are at s = -1 ± j√8
Comparing these values to c(t) in the same figure, we see that the real part of the
pole matches the exponential decay frequency of the sinusoid’s amplitude, while
the imaginary part of the pole matches the frequency of the sinusoidal oscillation.

2. We now compare the pole location to the plot.


The figure below shows a general, damped sinusoidal response for a second-
order system.

second-order step response components generated by complex poles

The transient response consists of an exponentially decaying amplitude generated


by the real part of the system pole times a sinusoidal waveform generated by the
imaginary part of the system pole. The time constant of the exponential decay is
equal to the reciprocal of the real part of the system pole. The value of the
imaginary part is the actual frequency of the sinusoid. This sinusoidal frequency is
given the name damped frequency of oscillation, ωd.

3. The steady-state response (unit step) was generated by the input pole located at
the origin. The type of response shown above is called an underdamped
response, one which approaches a steady-state value via a transient response
that is a damped oscillation.

Example: By inspection, write the form of the step response of the system below

author w chiyabu Page 72


Solution:

First we determine that the form of the forced response is a step.


Next we find the form of the natural response. Factoring the denominator of the
transfer function, we find the poles to be s = -5 ± j13.23. The real part, -5, is the
exponential frequency for the damping. It is also the reciprocal of the time constant
of the decay of the oscillations. The imaginary part, 13.23, is the radian frequency
for the sinusoidal oscillations.

Thus, c(t) = K1 + e-5t (K2 cos 13.23t + K3 sin 13.23t)


= K1 + K4e-5t (cos 13.23t – ϕ),

where ϕ = tan-1 (K3/K2); K4 = √(K22 + K23)


and c(t) is a constant plus an exponentially damped sinusoid.

Undamped Response

undamped system pole-zero plot response

For this response,

This function has a pole at the origin that comes from the unit step input and two
imaginary poles that come from the system. The input pole at the origin generates
the constant forced response, and the two system poles on the imaginary axis at ±j3
generate a sinusoidal natural response whose frequency is equal to the location of
the imaginary poles.
Hence, the output can be estimated as c(t) = K1 + K4 cos(3t – ϕ).
This type of response is called undamped. Note that the absence of a real part in
the pole pair corresponds to an exponential that does not decay. Mathematically, the
exponential is e-0t = 1.

author w chiyabu Page 73


Critically Damped Response

critically damped system pole-zero plot response

For this response,

This function has a pole at the origin that comes from the unit step input and two
multiple real poles that come from the system. The input pole at the origin
generates the constant forced response, and the two poles on the real axis at -3
generate a natural response consisting of an exponential and an exponential
multiplied by time, where the exponential frequency is equal to the location of the
real poles.

Hence, the output can be estimated as c(t) = K1 + K2e-3t + K3te-3t .

This type of response is called critically damped. Critically damped responses are
the fastest possible without the overshoot that is characteristic of the underdamped
response.

Summary of Natural Response

1. Overdamped Responses
Poles: Two real at -σ1, -σ2
Natural Response: Two exponentials with time constants equal to the reciprocal of
the pole locations, or c(t) = K1e-σ1t + K2e-σ2t

2. Underdamped Responses
Poles: Two complex at -σd ± jωd
Natural Response: Damped sinusoid with an exponential envelope whose time
constant is equal to the reciprocal of the pole’s real part. The radian frequency of
the sinusoid, the damped frequency of oscillation, is equal to the imaginary part
of the poles, or c(t) = Ae-σd*t cos(ωdt – ϕ)

3. Undamped Responses
Poles: Two imaginary at ±jω1
Natural Response: Undamped sinusoid with radian frequency equal to the
imaginary part of the poles, or c(t) = Acos(ω1t – ϕ)
author w chiyabu Page 74
4. Critically Damped Responses
Poles: Two real at -σ1
Natural Response: One term is an exponential whose time constant is equal to the
reciprocal of the pole location. Another term is the product of time, t, and an
exponential with time constant equal to the reciprocal of the pole location, or
c(t) = K1e-σ1t + K2te-σ1t

step responses for second-order system damping cases

Exercise
For each of the following transfer functions, write, by inspection, the general form of
the step response:

questions answers

author w chiyabu Page 75


7.4 The General Second-Order System

A second-order system can be generalised to establish quantitative specifications


defined in such a way that the response can be described to a designer without the
need for sketching the response. There are two physically meaningful specifications
for second-order systems that can be used to describe the characteristics of the
second-order transient response just as time constants describe the first-order
system response. The two quantities are called natural frequency and damping
ratio.

Natural Frequency, ωn
The natural frequency of a second-order system is the frequency of oscillation of the
system without damping. For example, the frequency of oscillation of a series RLC
circuit with the resistance shorted would be the natural frequency.

Damping Ratio, ζ
A second-order system’s underdamped step response is characterised by damped
oscillations. Our definition is derived from the need to quantitatively describe this
damped oscillation regardless of the time scale. Thus, a system whose transient
response goes through three cycles in a millisecond before reaching the steady state
would have the same measure as a system that went through three cycles in a
millennium before reaching the steady state. For example, the underdamped curve
has an associated measure that defines its shape. This measure remains the same
even if we change the time base from seconds to microseconds or to millennia.
A viable definition for this quantity is one that compares the exponential decay
frequency of the envelope to the natural frequency. This ratio is constant regardless
of the time scale of the response. Also, the reciprocal, which is proportional to the
ratio of the natural period to the exponential time constant, remains the same
regardless of the time base.

The damping ratio, ζ, is defined as

The general second-order system can be transformed to show the quantities ζ and
ωn.

Generalised second-order system

Without damping, the poles would be on the jω-axis, and the response would be an
undamped sinusoid. For the poles to be purely imaginary, a = 0.

author w chiyabu Page 76


Hence,

By definition, the natural frequency, ωn, is the frequency of oscillation of this system.
Since the poles of this system are on the jω-axis at ±j√b

ωn = √b or b = ω2n

Assuming an underdamped system, the complex poles have a real part, σ = -a/2.
The magnitude of this value is then the exponential decay frequency.

Hence,

From which a = 2ζωn

Modified general second-order transfer function:

Example: Given the transfer function below, find ζ and ωn

Solution:

Comparing to

ω2n = 36 from which ωn = 6 Also, 2ζωn = 4.2

Substituting the value of ωn, ζ = 0.35.

Thus natural frequency ωn = 6 damping ratio ζ = 0.35

author w chiyabu Page 77


Relationship of Natural frequency and Damping Ratio to Pole Location

From

The poles of the transfer function are given by

Second-Order System Responses as a function of Damping Ratio

author w chiyabu Page 78


Example: For each of the systems shown below, find the value of damping ratio and
report the kind of response expected.

(a) (b) (c)

Solution:

From and

Since a = 2ζωn and ωn = √b,

Thus for system (a) ζ = 1.155 which is overdamped, since ζ > 1


for system (b) ζ = 1 which is critically damped
for system (c) ζ = 0.894 which is thus underdamped, since ζ < 1

Exercise

For each of the following transfer functions


(1) Find the values of damping ratio ζ and natural frequency ωn
(2) characterize the nature of the response.

Answers:
a. damping ratio, ζ = 0.3, natural frequency ωn = 20 system is underdamped
b. damping ratio, ζ = 1.5, natural frequency ωn = 30 system is overdamped
c. damping ratio, ζ = 1, natural frequency ωn = 15 system is critically damped
d. damping ratio, ζ = 0, natural frequency ωn = 25 system is undamped

author w chiyabu Page 79


7.5 Underdamped Second-Order Systems

Consider the general second-order system:

Considering the step response for the above general second-order system, the
transform of the response, C(s), is the transform of the input times the transfer
function, or

where it is assumed that ζ < 1 (the underdamped case).

Expanding by partial fractions, yields

Taking the inverse Laplace transform produces

where φ tan-1(ζ / √(1 – ζ2)).

A plot of this response appears below for various values of ζ, plotted along a time
axis normalised to the natural frequency.

second-order underdamped responses for damping ratio values

author w chiyabu Page 80


The parameters associated with the underdamped response are rise time, peak time,
percent overshoot, and settling time.

Definitions

1. Rise Time, Tr.


The time required for the waveform to go from 0.1 of the final value to 0.9 of the
final value.

2. Peak Time, TP.


The time required to reach the first, or maximum, peak.

3. Percent Overshoot, %OS.


The amount that the waveform overshoots the steady-state, or final, value at the
peak time, expressed as a percentage of the steady-state value.

4. Settling Time, Ts.


The time required for the transient’s damped oscillations to reach and stay within
±2% of the steady-state value.

These definitions are valid for systems of order higher than 2, although analytical
expressions for these parameters cannot be found unless the response of the
higher-order system can be approximated as a second-order system.

Rise time, peak time, and settling time yield information about the speed of the
transient response. This information can help a designer determine if the speed and
the nature of the response do or do not degrade the performance of the system.

second-order underdamped response specifications

author w chiyabu Page 81


Evaluation of Peak Time, Tp
Peak Time, Tp is found by differentiating c(t) and finding the first zero crossing after
t = 0.
This task is simplified by “differentiating” in the frequency domain.
Assuming zero initial conditions and using

we get

Completing the squares in the denominator, we have

Setting the derivative equal to zero yields

or

Each value of n yields the time for local maxima or minima. Letting n = 0 yields t =
0, the first point on the curve that has zero slope. The first peak, which occurs at the
peak time, Tp, is found by letting n =1:

Evaluation of Percent Overshoot, %OS


The percent overshoot, %OS, is given by

The term cmax is found by evaluating c(t) at the peak time, c(Tp).

author w chiyabu Page 82


Using

and

we get

For the unit step, cfinal = 1

Thus percent overshoot %OS

The percent overshoot is a function only of the damping ratio, ζ.


Whereas the above expression allows one to find %OS given the damping ratio ζ,
the inverse of the equation allows one to solve for ζ given %OS.

The inverse is given by

Plot of Percent Overshoot versus Damping Ratio

author w chiyabu Page 83


Evaluation of Settling Time, Ts
In order to find the settling time, we must find the time for which

reaches and stays within ±2% of the steady-state value, cfinal. Settling Time is the
time it takes for the amplitude of the decaying sinusoid in c(t) to reach 0.02, or

This equation assumes that cos(ωn√(1 – ζ2t – φ)) = 1 at settling time.

The settling time, Ts is

Approximation for the Settling Time that will be used for all values of ζ

Evaluation of Rise Time, Tr


A precise analytical relationship between rise time and damping ratio, ζ, cannot be
found.
However, using a computer and

the rise time can be found. We first designate ωnt as the normalised time variable
and select a value for ζ. Using the computer, we solve for the values of ωnt that yield
c(t) = 0.9 and c(t) = 0.1. Subtracting the two values of ωnt yields the normalised
rise time, ωnTr, for that value of ζ.

Continuing in like fashion with other values of ζ, we obtain the results plotted in
below.

author w chiyabu Page 84


Normalised rise time versus damping ratio for a second-order underdamped response

Finding Tp , %OS, Ts , and Tr from a Transfer Function

Example: Given the transfer function

find Tp, %OS, Ts, and Tr

Solution:

ωn and ζ are calculated as 10 and 0.75, respectively. Now substitute ζ and ωn into

, , and

and find, respectively, that Tp = 0.475 second, %OS = 2.838, and Ts = 0.533
second. Using the table above, the normalised rise time is approximately 2.3
seconds. Dividing by ωn yields Tr = 0.23 second. This problem demonstrates that we
can find Tp, %OS, Ts, and Tr without the tedious task of taking an inverse Laplace
transform, plotting the output response, and taking measurements from the plot.

author w chiyabu Page 85


Relationship of Peak Time, Percent Overshoot, and Settling Time to the location of
the Poles that generate these characteristics

second-order underdamped responses for damping ratio values

The pole plot for a general, underdamped second-order system is shown below.

pole plot for an underdamped second-order system

From the Pythagorean Theorem, the radial distance from the origin to the pole is the
natural frequency, ωn, and the cosθ = ζ.

Comparing

and

with the pole location, we evaluate peak time and settling time in terms of the pole
location.

author w chiyabu Page 86


peak time settling time

where ωd is the imaginary part of the pole and is called the damped frequency of
oscillation, and σd is the magnitude of the real part of the pole and is the
exponential damping frequency.

The above relationships shows that:

1. Peak Time, Tp is inversely proportional to the imaginary part of the pole.


Since horizontal lines on the s-plane are lines of constant imaginary value, they
are also lines of constant peak time.

2. Settling Time, Ts is inversely proportional to the real part of the pole.


Since vertical lines on the s-plane are lines of constant real value, they are also
lines of constant settling time.

3. Since Damping Ratio, ζ = cosθ, radial lines are lines of constant damping ratio ζ.
Since percent overshoot is only a function of damping ratio ζ, radial lines are thus
lines of constant percent overshoot, %OS.

These concepts are depicted in the figure below, where lines of constant Peak
Time Tp, Settling Time Ts, and Percent Overshoot %OS are labelled on the s-
plane.

Lines of constant peak time, Tp, settling time, Ts, and percent overshoot, %OS.
Note: Ts2 < Ts1 ; Tp2 < Tp1; %OS1 < %OS2

The significance of the figure above can be better understood by examining the
actual step response of comparative systems.

author w chiyabu Page 87


Depicted below are the step responses as the poles are moved in a vertical direction,
keeping the real part the same.

1. When poles move with constant real part

Step responses of second-order underdamped systems as poles move with constant real part

As the poles move in a vertical direction, the frequency increases, but the envelope
remains the same since the real part of the pole is not changing. The figure shows a
constant exponential envelope, even though the sinusoidal response is changing
frequency. Since all curves fit under the same exponential decay curve, the settling
time is virtually the same for all waveforms. Note that as overshoot increases, the
rise time decreases.

2. When poles move with constant imaginary part

Step responses of second-order underdamped systems as poles move with constant imaginary part

Let us move the poles to the right or left. Since the imaginary part is now constant,
movement of the poles yields constant frequency over the range of variation of the
real part. As the poles move to the left, the response damps out more rapidly, while
the frequency remains the same.
Notice that the peak time is the same for all waveforms because the imaginary part
remains the same.

author w chiyabu Page 88


3. When poles move with constant damping ratio

Step responses of second-order underdamped systems as poles move with constant damping ratio

Moving the poles along a constant radial line yields does not change the percent
overshoot. Notice also that the responses look exactly alike, except for their speed.
The farther the poles are from the origin, the more rapid the response.

Finding Tp, %OS, and Ts from Pole Location

Example: Given the pole plot shown below, find ζ; ωn; Tp;%OS, and Ts

Solution:
The damping ratio is given by ζ = cosθ = cos[arctan(7/3)] = 0.394
The natural frequency, ωn, is the radial distance from the origin to the pole, or
ωn = √(72 + 32) = 7.616
The peak time is Tp = (π / ωd) = (π/7) = 0.449 s
The percent overshoot is

The settling time is Ts = 4/σd = 4/3 = 1.333 s

author w chiyabu Page 89


Transient Response through Component Design

Example: Given the system shown below, find J and D to yield 20% overshoot and a
settling time of 2 seconds for a step input of torque T(t)

rotational mechanical system

Solution:
First, the transfer function for the system is

From the transfer function ωn = √(K/J) and 2ζωn = D/J


But from the problem statement Ts = 2 = 4/(ζωn) or ζωn = 2
Hence, 2ζωn = 4 = D/J
Also ζ = 4/(2ζωn) = 2√(J/K)
A 20% overshoot implies ζ = 0.456
Hence, J/K = 0.052

From the problem statement, K = 5 N-m/rad.


Combining this value with 2ζωn = 4 = D/J and J/K = 0.052,
D = 1.04 N-m-s/rad, and J = 0.26 kg-m2

Exercise: Find ζ; ωn; Ts; Tp; Tr, and %OS for a system whose transfer function is

Solution:
ζ = 0.421; ωn = 19; Ts = 0.5 s; Tp = 0.182 s; Tr = 0.079 s; and %OS = 23.3%:

author w chiyabu Page 90


References
Katsuhiko, O. (2010) Modern Control Engineering, 5e, New Jersey: Prentice Hall

Raven, F. H. (1978) Automatic Control Engineering, New York: McGraw-Hill


Book Company

author w chiyabu Page 91

You might also like