Control and Electrical Circuits
Control and Electrical Circuits
Electric Circuits
Prepared by
Second Year
2019
Acknowledgments
This two-year curriculum was developed through a participatory and collaborative approach
between the Academic faculty staff affiliated to Egyptian Universities as Alexandria University, Ain
Shams University, Cairo University , Mansoura University, Al-Azhar University, Tanta University, Beni
Souef University , Port Said University, Suez Canal University and MTI University and the Ministry of
Health and Population(General Directorate of Technical Health Education (THE). The design of this
course draws on rich discussions through workshops. The outcome of the workshop was course
specification with Indented learning outcomes and the course contents, which served as a guide to
the initial design.
We would like to thank Prof.Sabah Al- Sharkawi the General Coordinator of General Directorate of
Technical Health Education, Dr. Azza Dosoky the Head of Central Administration of HR
Development, Dr. Seada Farghly the General Director of THE and all share persons working at
General Administration of the THE for their time and critical feedback during the development of
this course.
Special thanks to the Minister of Health and Population Dr. Hala Zayed and Former Minister of
Health Dr. Ahmed Emad Edin Rady for their decision to recognize and professionalize health
education by issuing a decree to develop and strengthen the technical health education curriculum
for pre-service training within the technical health institutes.
Contents
Course Description
This book presents an integrated lumped course targeting the electric circuits, and
control systems. The main objective is to provide the students the capability of dealing
with the general control problem with focus on practical electrical systems. Consequently,
the modeling and analysis of basic DC and AC electric circuits are presented in this book.
In addition, the basic electronic switching circuits used in industrial applications are also
presented. The control of electric systems which is the main core of this book is given a
significant focus.
The book consists of six chapters. The first chapter presents an introduction to
control systems. The objectives of this chapter include the identification of the
requirements of control systems. In addition, the chapter presents an overview of the
fundamentals of control systems used in industrial applications. The second chapter
presents the fundamental of DC electric circuits, and the basic theories for their
modeling, and analysis. The third chapter handles the single-phase AC electric circuits,
while the fourth chapter presents the electronic switches, and the single-phase switching
circuits used in industrial and power applications. Chapter 5 presents the modeling, and
representation of general control systems with focus on practical electric circuits, and
systems. The last chapter presents block diagrams representation of control systems, and
examples of the analysis of responses of control systems.
Core Knowledge
Core Skills
Course Contents
Assignments
Field Work
Interactive
Research
Lecture
Class
Lab
ID Topics
2 DC Electric Circuits 2 2 2
20 6 18 8
TOTAL HOURS (52)
ii
Chapter 1
Introduction to Control Systems
Objectives
The objectives of this chapter include the identification of the requirements of control
systems. In addition, the chapter presents an overview of the fundamentals of control systems
used in industrial applications.
Introduction
The control systems are formed by integrating elements. The function of these
elements is to maintain a process variable at a desired value or within a desired range of
values. Requiring the human operator to take all of the required corrective action manually is
impractical, or sometimes impossible, especially if a large number of indications must be
monitored simultaneously, and fast actions should be taken at appropriate times. For these reasons,
most systems are controlled automatically once they are operating under normal conditions.
Automatic controls greatly reduce the burden on the operator and make his or her job
manageable. Process variables requiring control in a system include, but are not limited to,
flow, level, temperature, pressure, voltage magnitude, and power flow. Some systems do not
require all of their process variables to be fully controlled.
Automatic control systems, neither replace nor relieve the operator of the
responsibility for maintaining the facility by performing fully automatic actions. The operation
of the control systems is periodically checked by the operators to verify proper operation. If a
control system fails, the operator must be able to take over and control the process manually.
In most cases, understanding how the control system works aids the operator in determining
the system is operating properly and which actions are required to maintain the system in a
safe condition. Automatic control systems as essential parts in many applications such as
various industries, transport, and household equipment.
James Watt’s centrifugal governor for the speed control of a steam engine, in the
eighteenth century.
In 1922, Minorsky worked on automatic controllers for steering ships, and showed how
stability could be determined from the differential equations describing the system.
In 1932 Nyquist developed a relatively simple procedure for determining the stability of
a closed-loop system.
In 1934 Hasen introduced the term “servomechanisms” for position control systems,
and designed the relay servomechanisms, capable of closely following a changing input.
1
During the 1940’s frequency-response methods were used to design linear feedback
control systems.
From the end of the 1940’s to early 1950’s, the root-locus method in control system
design was fully developed.
Since the late 1950’s, the emphasis in control system design concentrated on the
design of optimal control systems.
Since about 1960, modern control systems with multi-input-multi-output variables were
developed.
Digital control systems emerged in the late 1970’s.
Highly developed computer control systems started since the early 1980’s.
2
Automatic regulating systems. An automatic regulating system is a feedback control
system in which the reference input or the desired output is either constant or slowly
varying with time, and in which the primary task is to maintain the actual output at the
desired value despite the presence of disturbances.
Process control systems. An automatic regulating system in which the output is a
variable such as temperature, pressure, flow, or liquid level is called by a process
control system.
Controlled variable. A controlled variable is the process variable that is maintained at
a specified value or within a specified range. In the previous example, the storage tank
level is the controlled variable.
Manipulated variable. A manipulated variable is the process variable that is acted on
by the control system to maintain the controlled variable at the specified value or
within the specified range. In the previous example, the flow rate of the water
supplied to the tank is the manipulated variable.
An advantage of the closed-loop control system is that the use of feedback makes the
system response relatively insensitive to external disturbances or internal variations in system
parameters. From the point of view of stability, the open-loop control system is easier to
3
build since stability is not a major problem. On the other hand, stability is always a major
problem in the closed-loop control system since it may tend to overcorrect error which may
cause oscillations of constant or changing amplitude. A proper combination of open-loop and
closed-loop controls will usually less expensive and give satisfactory overall system
performances. Figs. 3, and 4 illustrate the open-loop, and closed-loop control of a thermal
system. The reader is required to explain the functions of these control options.
4
Functions of Automatic Control
In any automatic control system, the four basic functions that occur are:
1. Measurement
2. Comparison
3. Computation
4. Correction
In the water tank level control system in the example above, the level transmitter measures
the level within the tank. The level transmitter sends a signal representing the tank level to
the level control device, where it is compared to a desired tank level. The level control
device, then computes how far to open the supply valve to correct any difference between
actual and desired tank levels.
1. A measurement element
2. An error detection element
3. A final control element
Relationships between these elements and the functions they perform in an automatic control
system are shown in Fig. 5. The measuring element performs the measuring function by
sensing and evaluating the controlled variable. The error detection element first compares
the value of the controlled variable to the desired value, and then signals an error if a
deviation exists between the actual and desired values. The final control element responds to
the error signal by correcting the manipulated variable of the process.
5
Fig. 5: Relationships between functions, and elements in an automatic control system
6
Illustrative Examples of Control Systems
7
Fig. 8: Computer control system.
Control Loop Diagrams
A loop diagram is a "roadmap" that traces the process flow through the system and
designates variables that can disrupt the balance of the system. A block diagram is a
pictorial representation of the cause and effect relationship between the input and output of
a physical system. A block diagram provides a mean to easily identify the functional
relationships among the various components of a control system.
The simplest form of a block diagram is the block and arrows diagram. It consists of a
single block with one input and one output (Figure 9 (A)). The block normally contains the
name of the element (Figure 9(B)) or the symbol of a mathematical operation (Figure 9(C)) to
be performed on the input to obtain the desired output. Arrows identify the direction of
information or signal flow.
8
Fig. 10: Summation points
9
Fig. 12: Feedback control system block diagram
Fig. 13: Lubrication oil cooler – the block diagram of the temperature control system
10
capacities. In Figure 13, the walls of the lube oil cooler oppose the transfer of heat from the
lube oil inside the tubes to the cooling water outside the tubes.
Transportation time is the time required to carry a change in a process variable from
one point to another in the process. If the temperature of the lube oil (Figure 13) is lowered
by increasing the cooling water flow rate, some time will elapse before the lube oil travels
from the lube oil cooler to the temperature transmitter. If the transmitter is moved farther
from the lube oil cooler, the transportation time will increase. This time lag is not just a
slowing down or retardation of a change; it is an actual time delay during which no change
occurs.
Oscillations describe this cyclic characteristic. There are three types of oscillations
that can occur in a control loop. They are decreasing amplitude, constant amplitude, and
increasing amplitude. Each is shown in Figure 14.
Decreasing amplitude (Figure 14(A)). These oscillations decrease in amplitude and
eventually stop with a control system that opposes the change in the controlled variable. This
is the condition desired in an automatic control system.
Constant amplitude (Figure 14(B)). The action of the controller sustains oscillations of
the controlled variable. The controlled variable will never reach a stable condition; therefore,
this condition is not desired.
Increasing the amplitude (Figure 14(C)). The control system not only sustains
oscillations, but also increases them. The control element has reached its full travel limits
and causes the process to go out of control.
11
Two position Control Systems
A two position controller is the simplest type of controller. A controller is a device that
generates an output signal based on the input signal it receives. The input signal is actually an
error signal, which is the difference between the measured variable and the desired value, or
setpoint; Figure 15.
This input error signal represents the amount of deviation between where the process system
is actually operating and where the process system is desired to be operating. The controller
provides an output signal to the final control element, which adjusts the process system to
reduce this deviation. The characteristic of this output signal is dependent on the type, or
mode, of the controller. This part describes the simplest type of controller, which is the two-
position, or ON-OFF, mode controller.
A two position controller is a device that has two operating conditions: completely on
or completely off. Figure 16 shows the input to output, the characteristic waveform for a two
position controller that switches from its "OFF" state to its "ON" state when the measured
variable increases above the setpoint. Conversely, it switches from its "ON" state to its "OFF"
state when the measured variable decreases below the setpoint. This device provides an
output determined by whether the error signal is above or below the setpoint. The magnitude
of the error signal is above or below the setpoint. The magnitude of the error signal past that
point is of no concern to the controller.
12
Fig. 17: Example of a two-position controller – volume of water in a tank
Valve Actuators
13
Fig. 18: Pneumatic actuator – Air to close, and spring to open
A rubber diaphragm separates the actuator housing into two air chambers. The upper chamber
receives supply air through an opening in the top of the housing. The bottom chamber
contains a spring that forces the diaphragm against the mechanical stops in the upper
chamber. Finally, a local indicator is connected to the stem to indicate the position of the
valve. The position of the valve is controlled by varying supply air pressure in the upper
chamber. This results in a varying force on the top of the diaphragm. Initially, with no supply
air, the spring forces the diaphragm upward against the mechanical stops and holds the valve
fully open. As supply air pressure is increased from zero, its force on top of the diaphragm
begins to overcome the opposing force of the spring. This causes the diaphragm to move
downward and the control valve to close. With increasing supply air pressure, the diaphragm
will continue to move downward and compress the spring until the control valve is fully
closed. Conversely, if supply air pressure is decreased, the spring will begin to force the
diaphragm upward and open the control valve. Additionally, if supply pressure is held
constant at some value between zero and maximum, the valve will position at an
intermediate position. Therefore, the valve can be positioned anywhere between fully open
and fully closed in response to changes in supply air pressure.
A positioner is a device that regulates the supply air pressure to a pneumatic actuator.
It does this by comparing the actuator’s demanded position with the control valve’s actual
position. The demanded position is transmitted by a pneumatic or electrical control signal
from a controller to the positioner. The pneumatic actuator in Figure 18 is shown in Figure 19
with a controller and positioner added.
14
Fig. 19: Pneumatic actuator with controller and positioner
The controller generates an output signal that represents the position demanded. This
signal is sent to the positioner. Externally, the positioner consists of an input connection of
the control signal, a supply air input connection, a supply air output connection, a supply air
vent connection, and a feedback linkage. Internally, it contains an intricate network of
electrical transducers, air lines, valves, linkages, and necessary adjustments. Other
positioners may also provide controls for local valve positioning and gauges to indicate supply
air pressure and control air pressure (for pneumatic controllers). From an operator’s
viewpoint, a description of complex internal workings of a positioner is not needed.
Therefore, this discussion will be limited to inputs to and outputs from the positioner.
In Figure 19, the controller responds to a deviation of a controlled variable from set
point and varies the control output signal accordingly to correct the deviation. The control
output signal is sent to the positioner, which responds by increasing or decreasing the supply
air to the actuator. Positioning of the actuator and control valve is fed back to the positioner
through the feedback link. When the valve has reached the position demanded by the
controller, the positioner stops the change in supply air pressure and holds the valve at the
new position. This, in turn, corrects the controlled variable’s deviation from setpoint.
For example, as the control signal increases, a valve inside the positioner admits more
supply air to the actuator. As a result, the control valve moves downward. The linkage
transmits the valve position information back to the positioner. This forms a small internal
feedback loop for the actuator. When the valve reaches the position that correlates to the
control signal, the linkage stops the supply air flow to the actuator. This causes the actuator
to stop. On the other hand, if the control signal decreases, another valve inside the positioner
opens and allows the supply air pressure to decrease by venting the supply air. This causes the
valve to move upward and open. When the valve has opened to the proper position, the
positioner stops venting air from the actuator and stops the movement of the control valve.
An important safety feature is provided by the spring in an actuator. It can be designed
to position a control valve in a safe position if a loss of supply air occurs. At a loss of supply
air, the actuator in Figure 19 will fail open. This type of arrangement is referred to as "air-to-
close, spring-to-open" or simply "fail-open." Some valves fail in the closed position. This type
of actuator is referred to as "air-to-open, spring-to-close" or "fail-closed." This "fail-safe"
concept is an important consideration in critical facility designs.
15
Hydraulic Actuators. Pneumatic actuators are normally used to control processes
requiring quick and accurate response, as they do not require a large amount of motive force.
However, when a large amount of force is required to operate a valve (for example, the main
steam system valves), hydraulic actuators are normally used. Although hydraulic actuators
come in many designs, piston types are most common.
A typical piston-type hydraulic actuator is shown in Figure 20. It consists of a cylinder,
piston, spring, hydraulic supply and return line, and stem. The piston slides vertically inside
the cylinder and separates the cylinder to two chambers. The upper chamber contains the
spring and the lower chamber contains hydraulic oil.
The hydraulic supply and return line is connected to the lower chamber and allows hydraulic
fluid to flow to and from the lower chamber of the actuator. The stem transmits the motion
of the piston to a valve.
Initially, with no hydraulic fluid pressure, the spring force holds the valve in the closed
position. As fluid enters the lower chamber, pressure in the chamber increases. This pressure
results in a force on the bottom of the piston opposite to the force caused by the spring.
When the hydraulic force is greater than the spring force, the piston begins to move upward,
the spring compresses, and the valve begins to open. As the hydraulic pressure increases, the
valve continues to open. Conversely, as hydraulic oil is drained from the cylinder, the
hydraulic force becomes less than the spring force, the piston moves downward, and the valve
closes. By regulating the amount of oil supplied or drained from the actuator, the valve can
be positioned between fully open and fully closed.
The principles of operation of a hydraulic actuator are like those of the pneumatic
actuator. Each uses some motive force to overcome the spring force to move the valve. Also,
hydraulic actuators can be designed to fail-open or fail-closed to provide a fail-safe feature.
Electric Solenoid Actuators. A typical electric solenoid actuator is shown in Figure 21.
It consists of a coil, armature, spring, and stem. The coil is connected to an external current
supply. The spring rests on the armature to force it downward. The armature moves vertically
inside the coil and transmits its motion through the stem to the valve.
16
Fig. 21: Electric solenoid actuator
When current flows through the coil, a magnetic field forms around the coil. The
magnetic field attracts the armature toward the center of the coil. As the armature moves
upward, the spring collapses and the valve opens. When the circuit is opened and current
stops flowing in the coil, the magnetic field collapses. This allows the spring to expand and
shut the valve.
A major advantage of solenoid actuators is their quick operation. Also, they are much
easier to install than pneumatic or hydraulic actuators. However, solenoid actuators have two
disadvantages. First, they have only two positions: fully open and fully closed. Second, they
don’t produce much force, so they usually only operate relatively small valves.
Electric Motor Actuators. Electric motor actuators vary widely in their design and
applications. Some electric motor actuators are designed to operate in only two positions
(fully open or fully closed). Other electric motors can be positioned between the two
positions. A typical electric motor actuator is shown in Figure 22. Its major parts include an
electric motor, clutch and gear box assembly, manual handwheel, and stem connected to a
valve.
17
Fig. 22: Electric motor actuator
The motor moves the stem through the gear assembly. The motor reverses its rotation
to either open or close the valve. The clutch and clutch lever disconnects the electric motor
from the gear assembly and allows the valve to be operated manually with the handwheel.
Most electric motor actuators are equipped with limit switches, torque limiters, or both. Limit
switches de-energize the electric motor when the valve has reached a specific position.
Torque limiters de-energize the electric motor when the amount of turning force has reached
a specified value. The turning force normally is greatest when the valve reaches the fully
open or fully closed position. This feature can also prevent damage to the actuator or valve if
the valve binds in an intermediate position.
Note:
The requirement of reasonable relative stability and that of steady-state accuracy tend
to be incompatible.
In designing control systems, we therefore find it necessary to make the most effective
compromise between these two requirements.
18
Design steps:
Analysis. By the analysis of a control system, we mean the investigation, under specific
conditions, of the performance of the system whose mathematical model is known.
Since any system is made up with components, analysis must start with a mathematical
description of each component. Once a mathematical model of the complete system
has been derived, the manner in which analysis is carried out is independent of
whether the physical system is pneumatic, electrical, mechanical … etc.
Design. To design a system means to find one which accomplishes a given task. In
general, the design procedure is not straightforward and will require trial-and-error
method.
Synthesis. By synthesis, we mean finding by a direct procedure a system that will
perform in a specified way. Usually such a procedure is entirely mathematical from the
start to the end of the design process. Synthesis procedures are now available for linear
networks and for optimal systems.
The basic approach to the design of a practical control system will necessarily involve
trial-and-error procedures.
After the mathematical design has been completed, the control engineer simulates the
model on a computer to test the behavior of the resulting system in response to various
signals and disturbances.
The system may be required to be redesigned.
A satisfactory result will lead to the production of the prototype physical system.
19
Chapter 2
DC Electric Circuits
Objective
To understand the basic laws of Dc electric circuit
To study different techniques to analysis of DC electric circuit
Introduction
A circuit consists of electrical elements connected together. Engineers use electric circuits
to solve problems that are important to modem society. In particular:
1. Electric circuits are used in the generation, transmission, and consumption of electric
power and energy.
2. Electric circuits are used in the encoding, decoding, storage, retrieval, transmission, and
processing of information.
20
Charge may flow in an electric circuit. Current is the time rate of change of charge past a
given point. Charge is the intrinsic property of matter responsible for electric phenomena.
The current through a specified area is defined by the electric charge passing through the
area per unit of time. Thus, q is defined as the charge expressed in coulombs (C).
Charge is the quantity of electricity responsible for electric phenomena. Then we can express
current as
I = dq / dt (1)
The unit of current is the ampere (A); an ampere is 1 coulomb per second. Current is the time
rate of flow of electric charge past a given point.
EX 1 :
Find the current in an element when the charge entering the element is
q = 12/ C
where t is the time in seconds.
Solution
Recall that the unit of charge is coulombs, C. Then the current, from Eq. 1, is
I = dq / dt = 12 A
where the unit of current is amperes, A.
If the charge q is known, the current i is readily found using Eq. 1. Alternatively, if the
current i is known, the charge q is readily calculated. Note that from Eq. 2, we obtain :
q= ∫ dx (2)
The basic variables in an electrical circuit are current and voltage. These variables
describe the flow of charge through the elements of a circuit and the energy required to
cause charge to flow. Figure 3 shows the notation we use to describe a voltage. There are
two parts to this notation: a value (perhaps represented by a variable name) and an assigned
direction. The value of a voltage may be positive or negative. direction of a voltage is given
by its polarities (+, - ) . As a matter of vocabulary, we say that a voltage exists across an
element. Figure 3 shows that there are two ways to label the voltage across an element. The
voltage vba is proportional to the work required to move a positive charge from terminal a to
terminal b. On the other hand, the voltage v ab is proportional to the work required to move a
positive charge from terminal b to terminal a. We sometimes read v ba as “the voltage at
21
terminal b with respect to terminal a.” Similarly, v ab can be read as ‘the voltage at terminal a
with respect to terminal b.” Alternatively, we sometimes say that vba is the voltage drop
from terminal a to terminal b. The voltages v ab and vba are similar but different. They have
the same magnitude but different polarities. This means that
vab = —vba (3)
When considering vba, terminal b is called the “+ terminal” and terminal a is called the
terminal.” On the other hand, when talking about v ab, terminal a is called the “ +
terminal” and terminal b is called the “ V terminal.” The voltage across an element is the
work (energy) required to move a unit positive charge from the “—“ terminal to the “+”
terminal. The unit of voltage is the volt, V.
The equation for the voltage across the element is
V = dw/dq (4)
where v is voltage, w is energy (or work), and q is charge. A charge of 1 coulomb delivers an
energy of 1 joule as it moves through a voltage of 1 volt.
+ Vab —
a b
—Vba +
From Eq. 6, we see that the power is simply the product of the voltage across an element
times the current through the element. The power has units of watts.
Two circuit variables are assigned to each element of a circuit: a voltage and a current.
Figure 4 shows that there are two different ways to arrange the direction of the current and
the polarity of the voltage. In Figure 4-a, the current enters the circuit element at the +
terminal of the voltage and exits at the - terminal. In contrast, in Figure 4-b, the current
enters the circuit element at the - terminal of the voltage and exits at the + terminal. First,
22
consider Figure 4-a. When the current enters the circuit element at the + terminal of the
voltage and exits at the - terminal, the voltage and current are said to “adhere to the passive
convention.” In the passive convention, the voltage pushes a positive charge in the direction
indicated by the current. Accordingly, the power calculated by multiplying the element
voltage by the element current
I + V —
a b
(a)
I
a b
— V +
(b)
Fig. 4 (a) The passive convention is used for element voltage and current
(b) The passive convention is not used.
p = vi (7)
is the power absorbed by the element. (This power is also called “the power received by the
element” and “the power dissipated by the element.” )
Next, consider Figure 4-b. Here the passive convention has not been used. Instead, the
current enters the circuit element at the - terminal of the voltage and exits at the + terminal.
In this case, the voltage pushes a positive charge in the direction opposite to the direction
indicated by the current. Accordingly, when the element voltage and current do not adhere to
the passive convention, the power calculated by multiplying the element voltage by the
element current is the power supplied by the element. The power absorbed by an element
and the power supplied by that same element are
related by
power absorbed = —power supplied (8)
Ex. 2
Let us consider the element shown in Figure 4-a when v = 4 V and i = 10 A. Find the power
absorbed by the element and the energy absorbed over a 10-s interval.
Solution
The power absorbed by the element is
23
p = vi = 4 • 10 = 40 W
The energy absorbed by the element is
W=∫ = 40 * 10 = 400 J
EX 3 :
Fig. 5 shows four circuit elements identified by the letters A, B, C and D
I. Which of the devices supply 12 W?
II. Which of the devices absorb 12 W?
III. What is the value of the power received by device B?
IV. What is the value of the power delivered by device B?
V. What is the value of the power delivered by device D?
Fig. 5
Answers: (a) B and C, (b) A and D, (c) -1 2 W, (d) 12 W, (e) -12 W
In fact, the law of conservation of energy must be obeyed in any electric circuit. For this
reason, the algebraic sum of power in a circuit, at any instant of time, must be zero. On other
hand it can be represented as the algebraic sum of power absorbed in a circuit should equal
the algebraic sum of power supplied in a circuit.
Basic Laws
If the current does not change with time, but remains constant, we call it a direct current
(dc). A direct current (dc) is a current that remains constant with time. In the next parts of
the chapter we will deal with DC circuits.
Circuit Elements
An element is the basic building block of a circuit. An electric circuit is simply an
interconnection of the elements. Circuit analysis is the process of determining voltages across
(or the currents through) the elements of the circuit. There are two types of elements found
in electric circuits: passive elements and active elements. An active element is capable of
generating energy while a passive element is not. Examples of passive elements are resistors,
capacitors, and inductors. Typical active elements include generators, batteries, and
24
operational amplifiers.
The most important active elements are voltage or current sources that generally deliver
power to the circuit connected to them. There are two kinds of sources: independent and
dependent sources.
Fig. 6 symbols for independent voltage sources (a) used for constant or time-varying
voltage (b) used for constant voltage (dc)
Resistance is the physical property of an element or device that impedes the flow of
current; it is represented by the symbol R. where A is the cross-sectional area, ρ the
resistivity, L the length, and v the voltage across the wire element. Ohm, defined the
25
constant resistance R as
Ohm’s law
Ohm's law, which related the voltage and current, was published in 1827 as :
V=I*R volt (10)
Ohm’s law states that the voltage v across a resistor is directly proportional to the current i
flowing through the resistor.
The unit of resistance R was named the ohm in honor of Ohm and is usually abbreviated by
the Ω (capital omega) symbol, where 1 Ω= 1 V/A.
When resistance value is approaching infinite it is called open circuit and when the resistance
value is approaching zero is called short circuit as shown in Fig. 9
26
A useful quantity in circuit analysis is the reciprocal of resistance R, known as conductance
and denoted by G. Conductance is the ability of an element to conduct electric current; it is
measured in mho or siemens (S).
Ex 4 :
An electric iron draws 2 A at 120 V. Find its resistance.
Solution:
From Ohm’s law,
R = v / I = 120 / 2 = 60 Ω
Ex 5
In the circuit shown in Fig. 10, calculate the current i, the conductance G, and the power p.
Fig. 10
Solution:
The voltage across the resistor is the same as the source voltage (30 V) because the resistor
and the voltage source are connected to the same pair of terminals. Hence, the current is
I = V/ R = 30 / (5 *103) = 6 mA
The conductance is
G = 1/ R = 1 / (5 * 103) = 0.2 mS
We can calculate the power in various ways
P= v * I = I2* R = V2 * G = 180 mW
NODES, BRANCHES, AND LOOPS
Since the elements of an electric circuit can be interconnected in several ways, we need
to understand some basic concepts of network topology. To differentiate between a circuit
and a network, we may regard a network as an interconnection of elements or devices,
whereas a circuit is a network providing one or more closed paths. The convention, when
addressing network topology, is to use the word network rather than circuit. We do this even
though the words network and circuit mean the same thing when used in this context. In
network topology, we study the properties relating to the placement of elements in the
network and the geometric configuration of the network. Such elements include branches,
27
nodes, and loops. A branch represents a single element such as a voltage source or a
resistor.
In other-words, a branch represents any two-terminal element. The circuit in Fig. 11 has
five branches, namely, the 10-V voltage source, the 2-A current source, and the three
resistors.
28
Independent loops or paths result in independent sets of equations. For example, the closed
path abca containing the 2 Ω resistor in Fig. 12 is a loop. Another loop is the closed path bcb
containing the 3 Ω resistor and the current source. Although one can identify six loops in Fig.
12, only three of them are independent.
As the next two definitions show, circuit topology is of great value to the study of voltages
and currents in an electric circuit. Two or more elements are in series if they are cascaded or
connected sequentially and consequently carry the same current. Two or more elements are
in parallel if they are connected to the same two nodes and consequently have the same
voltage across them.
Elements are in series when they are chain-connected or connected sequentially, end to
end. For example, two elements are in series if they share one common node and no other
element is connected to that common node. Elements in parallel are connected to the same
pair of terminals. Elements may be connected in a way that they are neither in series nor in
parallel. In the circuit shown in Fig. 11, the voltage source and the 5 Ω resistor are in series
because the same current will flow through them. The 2 Ω resistor, the 3 Ω resistor, and the
current source are in parallel because they are connected to the same two nodes (b and c)
and consequently have the same voltage across them. The 5 Ω and 2 Ω resistors are neither in
series nor in parallel with each other.
KIRCHHOFF’S LAWS
Ohm’s law by itself is not sufficient to analyze circuits. However, when it is coupled with
Kirchhoff’s two laws, we have a sufficient, powerful set of tools for analyzing a large variety
of electric circuits. Kirchhoff’s laws were first introduced in 1847 by the German physicist
Gustav Robert Kirchhoff (1824–1887). These laws are formally known as Kirchhoff’s current
law (KCL) and Kirchhoff’s voltage law (KVL). Kirchhoff’s first law is based on the law of
conservation of charge, which requires that the algebraic sum of charges within a system
cannot change. Kirchhoff’s current law (KCL) states that the algebraic sum of currents
entering a node (or a closed boundary) is zero. Mathematically, KCL implies that
where N is the number of branches connected to the node and in is the nth current entering (or
leaving) the node. By this law, currents entering a node may be regarded as positive, while
currents leaving the node may be taken as negative or vice versa. Consider the node in Fig.
13. Applying KCL gives
29
since currents i1, i3, and i4 are entering the node, while currents i2 and i5 are leaving it. By
rearranging the terms, we get
i 1 + i3 + i4 = i2 + i5 (12)
The sum of the currents entering a node is equal to the sum of the currents leaving the node.
Fig. 14 Current sources in parallel (a) original circuit (b) equivalent circuit
A simple application of KCL is combining current sources in parallel. The combined current is
the algebraic sum of the current supplied by the individual sources. For example, the current
sources shown in Fig. 14 (a) can be combined as in Fig. 14 (b). The combined or equivalent
current source can be found by applying KCL to node a.
I T + I2 = I 1 + I 3
30
I T = I1 − I 2 + I 3 (13)
A circuit cannot contain two different currents, I 1 and I2, in series, unless I1 = I2; otherwise
KCL will be violated. Kirchhoff’s second law is based on the principle of conservation of
energy:
Kirchhoff’s voltage law (KVL) states that the algebraic sum of all voltages around a closed
path(or loop) is zero. Expressed mathematically, KVL states that
Where m is the number of voltages in the loop (or the number of branches in the loop) and
Vm is the mth voltage. To illustrate KVL, consider the circuit in Fig. 15. The sign on each
voltage is the polarity of the terminal encountered first as we travel around the loop. We can
start with any branch and go around the loop either clockwise or counterclockwise. Suppose
we start with the voltage source and go clockwise around the loop as shown; then voltages
would be −v1,+v2,+v3,−v4, and +v5, in that order. For example, as we reach branch 3, the
positive terminal is met first; hence we have+v 3. For branch 4, we reach the negative
terminal first; hence, −v4. Thus, KVL yields
−v1 + v2 + v3 − v4 + v5 = 0 (14)
v2 + v3 + v5 = v1 + v4
31
This is an alternative form of KVL. Notice that if we had traveled counterclockwise, the result
would have been +v1, −v5, +v4, −v3, and −v2, which is the same as before except that the signs
are reversed. Hence, Eqs. (14) and (15) remain the same. When voltage sources are
connected in series, KVL can be applied to obtain the total voltage. The combined voltage is
the algebraic sum of the voltages of the individual sources. For example, for the voltage
sources shown in Fig. 16 (a), the combined or equivalent voltage source in Fig. 16 (b) is
obtained by applying KVL.
−Vab + V1 + V2 − V3 = 0
or
Vab = V1 + V2 − V3 (16)
To avoid violating KVL, a circuit cannot contain two different voltages V1 and V2 in
parallel unless V1 = V2.
Fig. 16 Voltage sources in series (a) original circuit (b) equivalent circuit
EX
32
Fig. 17
Solution:
To find v1 and v2, we apply Ohm’s law and Kirchhoff’s voltage law. Assume that current
“i” flows through the loop as shown in Fig. 17 (b). From Ohm’s law,
−20 + v1 − v2 = 0 (18)
−20 + 2i + 3i =0 or 5i = 20 ⇒ i = 4 A
v1 = 8 V, v2 = −12 V
Ex
Fig. 18
Answer: 12 V, −6 V.
Ex: Find the currents and voltages in the circuit shown in Fig. 19 (a).
Fig. 19
33
Solution:
Since the voltage and current of each resistor are related by Ohm’s law as shown, we
are really looking for three things: (v1, v2, v3) or (i1, i2, i3). At node a, KCL gives
i 1 − i2 − i3 = 0 (20)
−30 + v1 + v2 = 0
or
−v2 + v3 = 0 ⇒ v3 = v2 (22)
as expected since the two resistors are in parallel. We express v1 and v2 in terms of i1
and i2 as in Eq. (19). Equation (22) becomes
(30 − 3i2)/8 − i2 − i2 /2 = 0 or i2 = 2 A.
From the value of i2, we now use Eqs. (19) to (23) to obtain
i1 = 3 A, i3 = 1 A, v1 = 24 V, v2 = 6 V, v3 = 6 V
Ex
Find the currents and voltages in the circuit shown in Fig. 20.
34
Fig. 20
35
or
v = iReq (28)
Implying that the two resistors can be replaced by an equivalent resistor; that
is,
Thus, Fig. 21 can be replaced by the equivalent circuit in Fig. 22. The two circuits in Figs. 21
and 22 are equivalent because they exhibit the same voltage-current relationships at the
terminals a-b. An equivalent circuit such as the one in Fig. 22 is useful in simplifying the
analysis of a circuit. In general, the equivalent resistance of any number of resistors
connected in series is the sum of the individual resistances. For N resistors in series then,
Notice that the source voltage v is divided among the resistors in direct proportion to their
resistances; the larger the resistance, the larger the voltage drop. This is called the principle
of voltage division, and the circuit in Fig. 21 is called a voltage divider. In general, if a
voltage divider has N resistors in series with the source voltage v, (R1, R2,……….RN) , the nth
resistor (Rn ) will have a voltage drop of
36
Parallel Resistors
Consider the circuit in Fig. 23, where two resistors are connected in parallel and therefore
have the same voltage across them. From Ohm’s law,
v =i1R1 = i2R2
i = i1+ i2 (34)
We can extend the result in Eq. (36) to the general case of a circuit with N resistors in
parallel. The equivalent resistance is
37
This means that we may replace the circuit in Fig. 23 with that in Fig. 24.
Given the total current i entering node a in Fig. 23, how do we obtain current i1 and i2 ? and
We know that the equivalent resistor has the same voltage, or
V =i Req = (38)
Thus
i1= i2 = (39)
Which shows that the total current i is shared by the resistors in inverse proportion to their
resistances. This is known as the principle of current division, and the circuit in Fig. 23 is
known as a current divider. Notice that the larger current flows through the smaller
resistance.
38
Ex: Find Req for the circuit in Figure 25
Solution
Fig. 25
Fig. 26
39
We notice that the two 2 Ω resistors are in series so the equivalent resistance is
This 4 Ω resistor is in parallel with the 6 Ω in Fig. 26 (a) their equivalent resistance is
The circuit in Fig. 26 (a) is replaced with that in Fig. 26 (b). Hence the equivalent
resistance is
Fig. 27
Answer: 6 Ω
Fig. 28
40
The 6 Ω and 3 Ω are in parallel. Their combined resistance is
Thus we can replace the circuit in Figure 28 by that in Figure 29 (a). In Figure 29 (a), the 6 Ω
and 2 Ω are in parallel gives 2 Ω. This 2 Ω equivalent is series with 1 Ω to give combined
resistance of 3 Ω. Thus we replace circuit in Figure 29 (a) by that in Figure 29 (b). In Figure
29 (b)we combine the 2 Ω and 3 Ω are in parallel to get
41
Fig. 29
Ex: Find io and Vo in the circuit shown in Figure 30 (a). calculate the power dissipated in
the 3 Ω resistor.
Solution
Thus our circuit reduces to that in Figure 30 (b). we can get Vo by ohm’s law :
42
Fig. 30
Po = Vo io = 5.333 W
Nodal Analysis
Nodal analysis provides a general procedure for analyzing circuits using node voltages as the
circuit variables. Choosing node voltages instead of element voltages as circuit variables is
convenient and reduces the number of equations one must solve simultaneously. To simplify
matters, we shall assume in this section that circuits do not contain voltage sources. Circuits
that contain voltage sources will be analyzed in the next section. In nodal analysis, we are
interested in finding the node voltages. Given a circuit with n nodes without voltage sources,
the nodal analysis of the circuit involves taking the following three steps.
1. Select a node as the reference node. Assign voltages v1, v2, …., vn to the remaining
n-1 nodes. The voltages are referenced with respect to the reference node.
2. Apply KCL to each of the n-1 non-reference nodes. Use Ohm’s law to express the
branch currents in terms of node voltages.
43
We shall now explain and apply these three steps. The first step in nodal analysis is selecting
a node as the reference or datum node. The reference node is commonly called the ground
since it is assumed to have zero potential. A reference node is indicated by any of the three
symbols in Fig. 31. The type of ground in Fig. 31 (c) is called a chassis ground and is used in
devices where the case, enclosure, or chassis acts as a reference point for all circuits. When
the potential of the earth is used as reference, we use the earth ground in Fig. 31 (a) or
(b). We shall always use the symbol in Fig. 31 (b). Once we have selected a reference node,
we assign voltage designations to non-reference nodes. Consider, for example, the circuit in
Fig. 32 (a). Node 0 is the reference node (v=0). While nodes 1 and 2 are assigned voltages v1
and v2 respectively. Keep in mind that the node voltages are defined with respect to the
reference node. As illustrated in Fig. 32 (a), each node voltage is the voltage rise from the
reference node to the corresponding non-reference node or simply the voltage of that node
with respect to the reference node as the second step, we apply KCL to each non-reference
node in the circuit. To avoid putting too much information on the same circuit, the circuit in
Fig. 32 (a) is redrawn in Fig. 32 (b), where we now add i1, i2 and i3 as the currents through
resistors R1, R2 and R3 respectively. At node 1, applying KCL gives
I 1 = I2 + i 1 + i 2 (40)
At Node 2,
I2 +i2 = i3 (41)
Fig. 31 Common symbols for reference node (a) common ground (b) ground (c) chassis
ground
44
Fig. 32 Nodal analysis circuit
We now apply Ohm’s law to express the unknown currents i1, i2 and i3 in terms of node
voltages. The key idea to bear in mind is that, since resistance is a passive element, by the
passive sign convention, current must always flow from a higher potential to a lower
potential. Current flows from a higher potential to a lower potential in a resistor. We can
express this principle as
45
Then, by substitution in Eqs. 40 and 41
The third step in nodal analysis is to solve for the node voltages. If we apply KCL to n-1 non-
reference nodes, we obtain n-1 simultaneous equations. We solve Eqs. to obtain the node
voltages v1 and v2 using any standard method, such as the substitution method, the
elimination method, Cramer’s rule, or matrix inversion. To use either of the last two
methods, one must cast the simultaneous equations in matrix form.
Example: Calculate the node voltages in the circuit shown in Figure 33 (a)
Solution
Consider Figure 33 (b) where the circuit in Figure 33 (a) has been prepared for nodal analysis .
20 = V1 – V2 + 2 V1
Or
3 V1 – V2 = 20 (42)
Or
-3 V1 + 5 V2 = 60 (43)
46
By solving equations 42 and 43
V1 = 40/3 = 13.333 V
V2 = 20 V
Fig. 33
47
Example: Obtain the node voltages in the circuit of Fig. 34.
Fig. 34
■ CASE 1 If a voltage source is connected between the reference node and a non-reference
node, we simply set the voltage at the non-reference node equal to the voltage of the
voltage source. In Fig. 35, for example, V1 = 10 volt Thus, our analysis is somewhat simplified
by this knowledge of the voltage at this node.
■ CASE 2 If the voltage source (dependent or independent) is connected between two non-
reference nodes, the two non-reference nodes
48
form a generalized node or supernode; we apply both KCL and KVL to determine the node
voltages. A supernode is formed by enclosing a (dependent or independent) voltage source
connected between two non-reference nodes and any elements connected in parallel with it.
In Fig. 36, nodes 2 and 3 form a supernode. (We could have more than two nodes forming a
single supernode.
We analyze a circuit with supernodes using the same three steps mentioned in the previous
section except that the supernodes are treated differently. Why? Because an essential
component of nodal analysis is applying KCL, which requires knowing the current through
each element. There is no way of knowing the current through a voltage source in advance.
However, KCL must be satisfied at a supernode like any other node. Hence, at the supernode
in Fig. 35,
To apply Kirchhoff’s voltage law to the supernode in Fig. 35, we redraw the circuit as shown
in Fig. 36. Going around the loop in the clockwise direction gives
Fig. 36
49
Note the following properties of a supernode:
1. The voltage source inside the supernode provides a constraint equation needed to
solve for the node voltages.
Fig. 37
Mesh Analysis
Mesh analysis provides another general procedure for analyzing circuits, using mesh
currents as the circuit variables. Using mesh currents instead of element currents as
circuit variables is convenient and reduces the number of equations that must be solved
simultaneously. Recall that a loop is a closed path with no node passed more than once. A
mesh is a loop that does not contain any other loop within it.
In Fig. 38, for example, paths abefa and bcdeb are meshes, but path abcdefa is not a
mesh. The current through a mesh is known as mesh current. In mesh analysis, we are
50
interested in applying KVL to find the mesh currents in a given circuit. In this section, we
will apply mesh analysis to planar circuits that do not contain current sources. In the next
section, we will consider circuits with current sources. In the mesh analysis of a circuit
with n meshes, we take the following three steps. Steps to Determine Mesh Currents:
2. Apply KVL to each of the n meshes. Use Ohm’s law to express the voltages in terms of
the mesh currents.
To illustrate the steps, consider the circuit in Fig. 38. The first step requires that mesh
currents i1 and i2 are assigned to meshes 1 and 2. Although a mesh current may be
assigned to each mesh in an arbitrary direction, it is conventional to assume that each
mesh current flows clockwise. As the second step, we apply KVL to each mesh. Applying
KVL to mesh 1, we obtain
(44)
(45)
Note in the Eq. 44, that the coefficient of i1 is the sum of the resistances in the first
mesh, while the coefficient of i2 is the negative of the resistance common to meshes 1
and 2. Now observe that the same is true in Eq. (45). This can serve as a shortcut way of
writing the mesh equations. The third step is to solve for the mesh currents. Putting Eqs.
(44) and (45) in matrix form yields
51
Which can be solved to obtain the mesh currents i1 and i2 . Notice that the branch
currents are different from the mesh currents unless the mesh is isolated. To distinguish
between the two types of currents, we use i for a mesh current and I for a branch
current. The current elements I1, I2 and I3 are algebraic sums of the mesh currents. It is
evident from Fig. 38 that
Ex: Calculate the mesh currents i1 and i2 of the circuit of Fig. 40.
Fig. 40
Answer: i1 = 2 A, i2 = 0 A.
■ CASE 1 When a current source exists only in one mesh: Consider the circuit in Fig. 41, for
example. We set i2 = -5 A and write a mesh equation for the other mesh in the usual way;
that is,
52
Fig. 41
■ CASE 2 When a current source exists between two meshes: Consider the circuit in
Fig. 42 (a), for example. We create a supermesh by excluding the current source
and any elements connected in series with it, as shown in Fig. 42 (b). Thus, A
supermesh results when two meshes have a (dependent or independent) current
source in common.
Fig. 42 (a) Two meshes having a current source in common (b) a supermesh
As shown in Fig. 42 (b), we create a supermesh as the periphery of the two meshes and
treat it differently. (If a circuit has two or more supermeshes that intersect, they should
be combined to form a larger supermesh.) Why treat the supermesh differently? Because
mesh analysis applies KVL—which requires that we know the voltage across each branch—
and we do not know the voltage across a current source in advance. However, a
supermesh must satisfy KVL like any other mesh. Therefore, applying KVL to the
supermesh in Fig. 42 (b) gives
53
We apply KCL to a node in the branch where the two meshes intersect. Applying KCL to
node 0 in Fig. 42 (a) gives
i2 = i1 + 6
Then by solving the previous equations
i1 = -3.2 A
i2 = 2.8 A
Fig. 43
Superposition
If a circuit has two or more independent sources, one way to determine the value of a
specific variable (voltage or current) is to use nodal or mesh analysis. Another way is to
determine the contribution of each independent source to the variable and then add
them up. The latter approach is known as the superposition. The idea of superposition
54
rests on the linearity property. The superposition principle states that the voltage across
(or current through) an element in a linear circuit is the algebraic sum of the voltages
across (or currents through) that element due to each independent source acting alone.
The principle of superposition helps us to analyze a linear circuit with more than one
independent source by calculating the contribution of each independent source
separately. However, to apply the superposition principle, we must keep two things in
mind:
2. Dependent sources are left intact because they are controlled by circuit
variables. With these in mind, we apply the superposition principle in three
steps.
Steps to Apply Superposition Principle:
1. Turn off all independent sources except one source. Find the output (voltage or
current) due to that active source using the techniques
3. Find the total contribution by adding algebraically all the contributions due to
the independent sources.
it may very likely involve more work. If the circuit has three independent sources,
we may have to analyze three simpler circuits each providing the contribution
due to the respective individual source.
55
Example: Use the superposition theorem to find v in the circuit of Fig. 44.
Fig. 44
Solution:
V= v1 + v2
Where v1 and v2 are the contributions due to the 6-V voltage source and the 3-A current
source, respectively. To obtain v1 we set the current source to zero, as shown in Fig. 45
(a). Applying KVL to the loop in Fig. 45 (a) gives
Thus,
V1 = 4 i1 = 2 V
To get V2, we set the voltage source to zero as in Fig. 45 (b) using current division
Hence,
V2 = 4 i3 = 8 V
And we find
V =V1 + V 2 = 2 + 8 = 10 V
56
Fig. 45
Example: Using the superposition theorem, find in the circuit of Fig. 46.
Fig. 46
Answer: 6 V.
Thevenin’s Theorem
It often occurs in practice that a particular element in a circuit is variable (usually called
the load) while other elements are fixed. As a typical example, a household outlet
terminal may be connected to different appliances constituting a variable load. Each time
the variable element is changed, the entire circuit has to be analyzed all over again. To
avoid this problem, Thevenin’s theorem provides a technique by which the fixed part of
the circuit is replaced by an equivalent circuit. According to Thevenin’s theorem, the
linear circuit in Fig. 47 (a) can be replaced by that in Fig. 47 (b). (The load in Fig. 47 may
be a single resistor or another circuit.) The circuit to the left of the terminals a-b in Fig.
57
47 (b) is known as the Thevenin equivalent circuit; it was developed in 1883 by M. Leon
Thevenin (1857–1926), a French telegraph engineer.
Fig. 47 Replacing a linear two terminal circuit by its Thevenin equivalent (a)
original (b) the Thevenin equivalent
Our major concern right now is how to find the Thevenin equivalent voltage Vth and
resistance Rth . To do so, suppose the two circuits in Fig. 47 are equivalent. Two circuits
are said to be equivalent if they have the same voltage-current relation at their
terminals. Let us find out what will make the two circuits in Fig. 47 equivalent. If the
terminals are made open-circuited (by removing the load), no current flows, so that the
open-circuit voltage across the terminals in Fig. 47 (a) must be equal to the voltage
source in Fig. 47 (b), since the two circuits are equivalent. Thus is the open-circuit
voltage across the terminals as shown in Fig. 48 (a); that is,
vth = voc
Again, with the load disconnected and terminals open circuited, we turn off all
independent sources. The input resistance (or equivalent resistance) of the dead circuit
58
at the terminals a-b in Fig. 47 (a) must be equal to Rth in Fig. 47 (b) because the two
circuits are equivalent. Thus, Rth is the input resistance at the terminals when the
independent sources are turned off, as shown in Fig. 48(b); that is,
Rth = Rin
If the network has no dependent sources, we turn off all independent sources. Rth is the
input resistance of the network looking between terminals a and b, as shown in Fig. 48
(b).
EX
Find the Thevenin equivalent circuit of the circuit shown in Fig. 49, to the left of the
terminals a-b Then find the current through RL= 6 Ω, 16 Ω and 36 Ω
Fig. 49
Solution:
We find Rth by turning off the 32-V voltage source (replacing it with a short circuit) and
the 2-A current source (replacing it with an open circuit). The circuit becomes what is
shown in Fig. 50 (a).
Thus,
59
Fig. 50
To find VTH, Applying mesh analysis to the two loops in Fig. 50 (b) we obtain
i2 = -2 A
Then
i1 = 0.5 A
VTH = 12 (i1 –i2) = 30 V
IL = 30 / (RL + RTH) = 30 / (RL + 4)
When RL = 6,
IL = 30 / 10 =3 A
When RL = 16
IL =30/20 = 1.5 A
Finally at RL = 36,
IL = 0.75 A
Fig. 60
60
Note that in Fig. 61 the Maximum power is transferred to the load when the load
resistance equals the Thevenin resistance as seen from the load (RL= RTh). It can
be proved by getting Pload as a function of RL from Fig. 62. Then differentiate the Pload
equation with respect to RL. Finally, put the differentiation output equal to zero.
61
EX : Using Thevenin’s theorem, find the equivalent circuit to the left of the terminals in
the circuit of Fig. 63. Then find I.
Fig. 63
62
Chapter 3
Single-phase AC electric circuits
Objective
To understand the basic principles of Ac electric circuit.
To study different techniques to analysis of AC electric circuits.
Overview
Thus far our analysis has been limited for the most part to dc circuits: those circuits excited
by constant or time-invariant sources. We have restricted the forcing function to dc sources
for the sake of simplicity, for pedagogic reasons, and also for historic reasons. Historically, dc
sources were the main means of providing electric power up until the late 1800s. At the end
of that century, the battle of direct current versus alternating current began. Both had their
advocates among the electrical engineers of the time. Because ac is more efficient and
economical to transmit over long distances, ac systems ended up the winner. Thus, it is in
keeping with the historical sequence of events that we considered dc sources first. We now
begin the analysis of circuits in which the source voltage or current is time-varying. In this
chapter, we are particularly interested in sinusoidally time-varying excitation, or simply,
excitation by a sinusoid. A sinusoid is a signal that has the form of the sine or cosine function.
We begin with a basic discussion of sinusoids and phasors. We then introduce the concepts of
impedance and admittance. The basic circuit laws, Kirchhoff’s and Ohm’s, introduced for dc
circuits, will be applied to ac circuits. Consider the sinusoidal voltage
V (t) = Vm sin ωt (1)
Where the sinusoid is shown in Fig. 1(a) as a function of its argument ω and in Fig. 1 as a
function of time. It is evident that the sinusoid repeats itself every T seconds; thus, T is
called the period of the sinusoid. From the two plots in Fig. 1, we observe that
T=2π/ω (2)
The fact that V(t) repeats itself every T seconds is shown by replacing t by t+T in
Eq. (1). We get
63
That is V, has the same value at t+T as it does at t and v(t) is said to be periodic. As
mentioned, the period T of the periodic function is the time of one complete cycle or the
number of seconds per cycle. The reciprocal of this quantity is the number of cycles per
second, known as the cyclic frequency f of the sinusoid. Thus,
f= 1 / T (3)
From Eqs. (2) and (3), it is clear that
ω= 2 π f (4)
While ω is in radians per second (rad/s), f is in hertz (Hz). Let us now consider a more general
expression for the sinusoid,
V(t) = Vm sin(ωt + φ) (5)
Where (ωt + φ) is the argument and φ is the phase. Both argument and phase can be in
radians or degrees. Let us examine the two sinusoids shown in Fig. 2.
64
A sinusoid can be expressed in either sine or cosine form. When comparing two sinusoids, it is
expedient to express both as either sine or cosine with positive amplitudes. Using the
following relationships, we can transform a sinusoid from sine form to cosine form or vice
versa.
Sin (ωt ± 180°) = sin ωt
cos (ωt ±180°) = cos ωt
sin (ωt ±90°) = cos ωt
cos (ωt ±90°) = sin ωt
Example: Calculate the phase angle between V 1 = 10 cos (ωt + 50°) and
V2 = sin (ωt - 10°)
Solution
Thus
The phase difference between V1 and V2 is 30°
Phasors
Sinusoids are easily expressed in terms of phasors, which are more convenient to work with
than sine and cosine functions. A phasor is a complex number that represents the amplitude
and phase of a sinusoid. Phasors provide a simple means of analyzing linear circuits excited by
sinusoidal sources; solutions of such circuits would be intractable otherwise. The notion of
solving Ac circuits using phasors was first introduced by Charles Steinmetz in 1893. Before we
completely define phasors and apply them to circuit analysis, we need to be thoroughly
familiar with complex numbers.
A complex number z can be written in rectangular form as
Z=x+jy (7)
Where j=√-1; x is the real part of z; y is the imaginary part of z. In this context, the variables
x and y do not represent a location as in two-dimensional vector analysis but rather the real
and imaginary parts of z in the complex plane. Nevertheless, we note that there are some
resemblances between manipulating complex numbers and manipulating two-dimensional
65
vectors. The complex number z can also be written in polar or exponential form as
z = rԼφ = re jφ
Where r is the magnitude of z, and φ is the phase of z. We notice that z can be represented in
three ways:
Fig. 3 Representation of Vejωt (a)sinor rotating counterclockwise (b) its projection on the real
axis
Equation (12) states that to obtain the sinusoid corresponding to a given phasor V, multiply
the phasor by the time factor ejωt and take the real part. As a complex quantity, a phasor may
be expressed in rectangular form, polar form, or exponential form. Since a phasor has
magnitude and phase (“direction”), it behaves as a vector and is printed in boldface.
Equations (10) through (12) reveal that to get the phasor corresponding to a sinusoid, we first
66
express the sinusoid in the cosine form so that the sinusoid can be written as the real part of
a complex number. Then we take out the time factor ejωt and whatever is left is the phasor
corresponding to the sinusoid. By suppressing the time factor, we transform the sinusoid from
the time domain to the phasor domain. This transformation is summarized as follows:
This shows that the derivative of V(t) is transformed to the phasor domain as jωV
Solution
67
Phasor Relationships for Circuit Elements
Now that we know how to represent a voltage or current in the phasor or frequency domain,
one may legitimately ask how we apply this to circuits involving the passive elements R, L,
and C. What we need to do is to transform the voltage-current relationship from the time
domain to the frequency domain for each element. Again, we will assume the passive sign
convention. We begin with the resistor. If the current through a resistor R is i= Im cos(ωt+ φ)
the voltage across it is given by Ohm’s law as
V =I R = R Im cos(ωt+ φ)
The phasor form of voltage and current
V =R Im Լφ (14)
I = Im Լφ (15)
showing that the voltage-current relation for the resistor in the phasor domain continues to
be Ohm’s law, as in the time domain. Figure 4 illustrates the voltage-current relations of a
resistor. We should note from Eqs. (14) and (15) that voltage and current are in phase, as
illustrated in the phasor diagram in Fig. 5.
Fig. 4 Voltage current relation for a resistor in (a) time domain (b) frequency domain
For the inductor L, assume the current through it is i= Im cos(ωt+ φ) The voltage across the
inductor is
V = L di/dt =
=
Then
V=
But I = Im Լφ
Thus
V = j ω L I where =j
Figure 6 illustrates the voltage-current relations of an inductor.
68
Fig. 6 voltage current relations for an inductor in the (a) time domain (b) frequency domain
Showing that the voltage has a magnitude ω L Im of and a phase of φ + 90° The voltage and
current are 90° out of phase. Specifically, the current lags the voltage by 90°.
For the capacitor C, assume the voltage across it is V = Vm cos (ω t + φ). The current through
the capacitor is
I= C dv/dt
Similarly we can get that,
I=jωCV
Then
V = I / jωc
Showing that the current and voltage are 90° out of phase. To be specific, the current leads
the voltage by 90°. Figure 7 illustrates the voltage-current relations of a capacitor.
Fig. 7 voltage current relations for a capacitor in the (a) time domain (b) frequency domain
Table 2 summary
69
Example: The voltage V = 12 cos (60t +45°) is applied to a 0.1 H inductor Find the current
through the inductor.
Solution
V = jωLI
ω = 60 rad /s
V = 12 Լ45
I = V / jωl = 12 Լ 45 / (j60*0.1) = 2 Լ-45 A
iI(t) = 2 cos (60t -45°) A
Example: The voltage V = 10 cos (100t +30°) is applied to a 50μF capacitor. Find the current
through the capacitor.
Answer: 50 cos (100t +120°) mA
Or
Where G is called the conductance and B is called the susceptance. Admittance, conductance,
and susceptance are all expressed in the unit of siemens (or mhos). Table 3 shows the
summary.
70
Table 3 Impedance and admittance of passive elements
Fig. 8
Fig. 9
Answer: 8.944 sin(10t + 93.43° ) V , 4.472 sin(10t + 3.43° ) A
71
If we let I1, I2, ……..,In are the phasors form of current leaving or entering a closed surface
then
I1+I2+………….IN=0
Consider the N series-connected impedances shown in Fig. 10. The same current “I” flows
through the impedances.
Showing that the total or equivalent impedance of series-connected impedances is the sum of
the individual impedances. This is similar to the series connection of resistances. If N =2 as
shown in Fig. 11, the current through the impedances Is
72
Fig.12 ” N” impedance in series
This indicates that the equivalent admittance of a parallel connection of admittances is the
sum of the individual admittances. When as N=2 shown in Fig. 13, the currents in the
impedances are
Example: Find the input impedance in Fig. 14. Assume that the circuit operates at ω
= 50 rad /s.
73
Fig. 14
Solution
Fig. 15
74
Example: calculate Vo in the circuit of Figure 16
Fig. 16
Analyze AC Circuits
Analyzing Ac circuits usually requires three steps. Steps to Analyze AC Circuits:
1. Transform the circuit to the phasor or frequency domain.
2. Solve the problem using circuit techniques (nodal analysis, mesh analysis, superposition,
etc.).
3. Transform the resulting phasor to the time domain.
Note, Step 1 is not necessary if the problem is specified in the frequency domain. In step 2,
the analysis is performed in the same manner as dc circuit analysis except that complex
numbers are involved.
Nodal Analysis
The basis of nodal analysis is Kirchhoff’s current law. Since KCL is valid for phasors, we can
analyze ac circuits by nodal analysis. The following examples illustrate this.
Example: compute V1 and V2 in the circuit of Figure 17
Fig. 17
Solution
Applying KCL at supernode shown in Figure 18 (Nodes 1 and 2)
75
Fig. 18
But a voltage source is connected between nodes 1 and 2 so that
Then
Thus
Fig. 19
Mesh Analysis
Kirchhoff’s voltage law (KVL) forms the basis of mesh analysis. The validity of KVL for ac
circuits was shown before and is illustrated in the following examples. Keep in mind that the
very nature of using mesh analysis is that it is to be applied to planar circuits.
Example Determine current in the circuit of Fig. 20 using mesh analysis.
Solution:
Applying KVL to mesh 1, we obtain
76
(8 + j10 -j2) I1 - (-j2) I2 - j10I3 =0
Fig. 20
For mesh 3, I3 = 5 A
By sub by I3 in the previous equations
77
Fig. 21
Answer: 3.582 Լ65.45° A
Fig. 22
Solution
As shown in Figure 24, meshes 3 and 4 form a supermesh
For mesh 1, KVL gives
– (-j2) I2 – 8 I3 = 0
For mesh 2,
I2 =
For the supermesh
+ (-j5) I2 – 8 I1 = 0
From current source between meshes 3 and 4
I4 = I 3 + 4
Fig. 24
By solving the previous equations we get
I1= 3.618 Լ 274.5° A
V o= - ) = - 7.2134 – j 6.568 V
78
Superposition Theorem
Since ac circuits are linear, the superposition theorem applies to Ac circuits the same way it
applies to dc circuits. The theorem becomes important if the circuit has sources operating at
different frequencies. In this case, since the impedances depend on frequency, we must have
a different frequency domain circuit for each frequency. The total response must be obtained
by adding the individual responses in the time domain. It is incorrect to try to add the
responses in the phasor or frequency domain. Why? Because the exponential factor ejωt is
implicit in sinusoidal analysis, and that factor would change for every angular frequency ω. It
would therefore not make sense to add responses at different frequencies in the phasor
domain. Thus, when a circuit has sources operating at different frequencies, one must add
the responses due to the individual frequencies in the time domain.
Example: Find Vo in Figure 24 using the superposition theorem.
Fig. 24
Solution:
Since the circuit operates at three different frequencies ω=0 for the dc voltage source), one
way to obtain a solution is to use superposition, which breaks the problem into single-
frequency problems. So we let
Vo = V1 + V2 + V3
where V1 is due to the 5-V dc voltage source, V2 is due to the 10 cos2t voltage source, and
V3 is due to 2 sin 5t A current source. To find V1, we set to zero all sources except the 5-V dc
source. We recall that at steady state, a capacitor is an open circuit to dc while an inductor is
a short circuit to dc. There is an alternative way of looking at this. Since ω =0, jωL =0, 1/jωc
= either way, the equivalent circuit is as shown in Fig. 25(a).
By voltage division,
To find V2, we set to zero both the 5-V source and 2 sin 5t current source and transform the
circuit to the frequency domain
79
Fig. 25
By voltage division,
In time domain
To obtain V3 we set the voltage sources to zero and transform what is left to the frequency
domain
By current divider
In time domain
Thus
80
replaced by a voltage source in series with an impedance.
Fig. 27
Solution
We find ZTH by setting the voltage source to zero. As shown in Fig. 28(a), the 8 Ω resistance is
now in parallel with the – j 6 reactance, so that their combination gives
To find VTH consider the circuit in Fig. 28(b). Currents I1 and I2 are obtained as
81
Fig. 28
Example: Find the Thevenin equivalent at terminals a-b of the circuit in Fig. 29
Fig. 29
82
Chapter 4
Power electronic switching circuits
Objective
Overview of the Power electronic control of electric circuits
Understand the basics of Single-phase uncontrolled rectification circuits
Understand the basics of Single-phase controlled rectification circuits
Understand the basics of Single-phase inversion circuits
83
power electronics are becoming attractive for applications in motion control by replacing the
earlier electromechanical and electronic systems. Applications in power transmission and
renewable energy include high-voltage dc (VHDC) converter stations, flexible ac transmission
system (FACTS), static var compensators, and energy storage. In power distribution, these
include dc-to-ac conversion, dynamic filters, frequency conversion, and custom power
system. Almost all new electrical or electromechanical equipments, from household air
conditioners and computer power supplies to industrial motor controls, contain power
electronic circuits and/or systems. In order to keep up, working engineers involved in control
and conversion of power and energy into applications ranging from several hundred voltages
at a fraction of an ampere for display devices to about 10,000 V at high-voltage dc
transmission should have a working knowledge of power electronics.
DEFINITION Power electronics involves the study of electronic circuits intended to control
the flow of electrical energy. These circuits handle power flow at levels much higher than the
individual device ratings. Rectifiers are probably the most familiar examples of circuits that
meet this definition. Inverters (a general term for dc–ac converters) and dc–dc converters for
power supplies are also common applications. As shown in Fig. 1, power electronics
represents a median point at which the topics of energy systems, electronics, and control
converge and combine. Any useful circuit design for an energy application must address issues
of both devices and control, as well as of the energy itself. Among the unique aspects of
power electronics are its emphasis on large semiconductor devices, the application of
magnetic devices for energy storage, special control methods that must be applied to
nonlinear systems, and its fundamental place as a central component of today’s energy
systems and alternative resources. In any study of electrical engineering, power electronics
must be placed on a level with digital, analog, and radio-frequency electronics to reflect the
distinctive design methods and unique challenges.
84
Fig. 1 Control, energy and power electronics are interrelated
The history of power electronics has been closely allied with advances in electronic
devices that provide the capability to handle high power levels. Since about 1990, devices
have become so capable that a transition from a “device-driven” field to an “applications-
driven” field continues. This transition has been based on two factors: (1) advanced
semiconductors with suitable power ratings exist for almost every application of wide
interest, and (2) the general push toward miniaturization is bringing advanced power
electronics into a growing variety of products. Although the devices continue to improve,
their development now tends to follow innovative applications.
Key Characteristics
All power electronic circuits manage the flow of electrical energy between an electrical
source and a load. The parts in a circuit must direct electrical flows, not impede them. A
general power conversion system is shown in Fig. 2. The function of the power converter in
the middle is to control the energy flow between a source and a load. For our purposes, the
power converter will be implemented with a power electronic circuit. Because a power
converter appears between a source and a load, any energy used within the converter is lost
to the overall system. A crucial point emerges: to build a power converter, we should
consider only lossless components. A realistic converter design must approach 100%
efficiency. A power converter connected between a source and a load also affects system
reliability. If the energy source is perfectly reliable (it is available all the time), then a failure
in the converter affects the user (the load) just as if the energy source had failed. An
unreliable power converter creates an unreliable system. To put this in perspective, consider
that a typical American household loses electric power only a few minutes a year. Energy is
available 99.999% of the time. A converter must be better than this to prevent system
85
degradation. An ideal converter implementation will not suffer any failures over its
application lifetime. Extreme high reliability can be a more difficult objective than high
efficiency.
86
Fig. 3 A basic power electronic system
Diode as a Switch
Among all the static switching devices used in power electronics (PE), the power diode is
perhaps the simplest. Its circuit symbol is shown in Fig. 4. It is a two terminal device, and
terminal A is known as the anode whereas terminal K is known as the cathode. If terminal A
experiences a higher potential compared to terminal K, the device is said to be forward
biased and a current called forward current (IF ) will flow through the device in the direction
as shown. This causes a small voltage drop across the device (<1V), which in ideal condition is
usually ignored. On the contrary, when a diode is reverse biased, it does not conduct and a
practical diode does experience a small current flowing in the reverse direction called the
leakage current. Both the forward voltage drop and the leakage current are ignored in an
ideal diode. Usually in PE applications a diode is considered to be an ideal static switch. The
characteristics of a practical diode show a departure from the ideals of zero forward and
infinite reverse impedance, as shown in Fig. 5-a. In the forward direction, a potential barrier
associated with the distribution of charges in the vicinity of the junction, together with other
effects, leads to a voltage drop. This, in the case of silicon, is in the range of 1V for currents
in the normal range. In reverse, within the normal operating range of voltage, a very small
current flow which is largely independent of the voltage. For practical purposes, the static
characteristic is often represented by Fig. 5-b. In the Figure, the forward characteristic is
expressed as a threshold voltage Vo and a linear incremental or slope resistance, r.
87
The reverse characteristic remains the same over the range of possible leakage currents
irrespective of voltage within the normal working range.
Fig. 4 power diode (a) symbol; (b) and (c) types of packing
Properties of PN Junction
From the forward and reverse biased condition characteristics, one can notice that when
the diode is forward biased, current rises rapidly as the voltage is increased. Current in the
reverse biased region is significantly small until the breakdown voltage of the diode is
reached. Once the applied voltage is over this limit, the current will increase rapidly to a very
88
high value limited only by an external resistance.
DC diode parameters:
The most important parameters are the followings:
• Forward voltage, VF is the voltage drop of a diode across A and K at a defined current
level when it is forward biased.
• Breakdown voltage, VB is the voltage drop across the diode at a defined current level
when it is beyond reverse biased level. This is popularly known as avalanche.
• Reverse current IR is the current at a particular voltage, which is below the breakdown
voltage.
Single-phase uncontrolled rectification circuits
Electrical energy sources take the form of dc voltage sources at various values, sinusoidal
ac sources, polyphase sources, among others. A power electronic circuit might be asked to
transfer energy between two different dc voltage levels, between an AC source and a dc load,
or between sources at different frequencies. It might be used to adjust an output voltage or
power level, drive a nonlinear load, or control a load current. In this section, a few basic
converter arrangements are introduced, and energy conservation provides a tool for analysis.
EXAMPLE: Consider the circuit shown in Fig. 6. It contains an AC source, a switch, and a
resistive load. It is a simple but complete power electronic system.
89
Fig. 7 Input and output waveforms for the Example
and an RMS value equal to Vpeak/2. Since the output has nonzero dc voltage content, the
circuit can be used as an ac–dc converter. To make it more useful, a low-pass filter would be
added between the output and the load to smooth out the ac portion. This filters needs to be
lossless, and will be constructed from only inductors and capacitors. The circuit in the
previous Example acts as a half-wave uncontrolled rectifier with a resistive load. With the
hypothesized switch action, a diode (uncontrolled switch) can substitute for the ideal switch.
The example confirms that a simple switching circuit can perform power conversion functions.
But note that a diode is not, in general, the same as an ideal switch. A diode places
restrictions on the current direction, whereas a true switch would not. An ideal switch allows
control over whether it is on or off, whereas a diode’s operation is constrained by circuit
variables. Consider a second half-wave circuit, now with a series L–R load, shown in Fig. 8.
EXAMPLE: A series diode L–R circuit has ac voltage source input. This circuit operates much
differently than the half-wave rectifier with resistive load. A diode will be on if forward-
biased, and off if reverse-biased. In this circuit, when the diode is off, the current will be
zero.
90
Let us assume that the diode is initially off (this assumption is arbitrary, and we will check
it as the example is solved). If the diode is off, the diode current is i =0, and the voltage
across the diode will be Vac. The diode will become forward-biased when Vac becomes
positive. The diode will turn on when the input voltage makes a zero-crossing in the positive
direction. This allows us to establish initial conditions for the circuit: i(t 0) = 0, t0 = −π/(2ω).
The differential equation can be solved in a conventional way to give
91
diodes conducting at any given time. The direction of the current through the load is always
the same. This rectifier topology is known as the full bridge rectifier.
92
power electronic devices. Some thyristors are also controllable in switching from forward-
conduction back to a forward-blocking state. The particular design of a thyristor will
determine its controllability and often its application. Thyristors are typically used at the
highest energy levels in power conditioning circuits because they are designed to handle the
largest currents and voltages of any device technology (systems approximately with voltages
above 1 kV or currents above 100 A).
A thyristor used in some ac power circuits (50 or 60 Hz in commercial utilities or 400 Hz in
aircraft) to control ac power flow can be made to optimize internal power loss at the expense
of switching speed. These thyristors are called phase-control devices, because they are
generally turned from a forward blocking into a forward-conducting state at some specified
phase angle of the applied sinusoidal anode–cathode voltage waveform. A second class of
thyristors is used in association with dc sources or in converting ac power at one amplitude
and frequency into ac power at another amplitude and frequency, and must generally switch
on and off relatively quickly. A typical application for the second class of thyristors is in
converting a dc voltage or current into an Ac voltage or current. A circuit that performs this
operation is often called an inverter, and the associated thyristors used are referred to as
inverter thyristors. Figure 11 shows a conceptual view of a typical thyristor with the three p–n
junctions and the external electrodes labeled. Also shown in the Figure is the thyristor circuit
symbol used in electrical schematics. Figure 12 shows typical thyristor stud-mount and press-
pack packages.
93
Figure 12 typical thyristor stud-mount and press-pack packages.
Applications
The most important application of thyristors is for line frequency phase-controlled
rectifiers. This family includes several topologies, of which one of the most important is used
to construct HVDC transmission systems. A single-phase controlled rectifier is shown in Fig.
13. The use of thyristors instead of diodes allows the average output voltage to be controlled
by appropriate gating of the thyristors. If the gate signals to the thyristors were continuously
applied, the thyristors in Fig. 6.13 would behave as diodes. If no gate currents are supplied
they behave as open circuits. Gate current can be applied any time (phase delay) after the
forward voltage becomes positive. Using this phase-control feature, it is possible to produce
an average dc output voltage less than the average output voltage obtained from an
uncontrolled diode rectifier.
94
Single-phase controlled Half-wave Rectifier
The single-phase half-wave rectifier uses a single thyristor to control the load voltage as
shown in Fig. 14. The thyristor will conduct, on-state, when the voltage vT is positive and a
firing current pulse iG is applied to the gate terminal. The control of the load voltage is
performed by delaying the firing pulse by an angle α. The firing angle α is measured from the
position where a diode would naturally conduct. In case of Fig. 14, the angle α is measured
from the zero-crossing point of the supply voltage vs. The load in Fig. 14 is resistive and
therefore the current id has the same waveform of the load voltage. The thyristor goes to the
non-conducting condition, off-state, when the load voltage, and consequently the current,
reaches a negative value. The load average voltage is given by
Vdα = ∫ (6)
Where Vmax is the supply peak voltage. Hence, it can be seen from Eq. (6) that changing the
firing angle α controls both the load average voltage and the amount of transferred power.
Figure 15-a shows the rectifier waveforms for an R–L load.
When the thyristor is turned on, the voltage across the inductance is
VL = Vs – VR = L did/dt (7)
Where VR is the voltage in the resistance R, given by VR = R*id . If Vs − VR > 0, from Eq. (7)
holds that the load current increases its value. On the other hand, i d decreases its value when
Vs− VR < 0. The load current is given by
95
id(ωt) = ∫ (8)
Graphically, Eq. (8) means that the load current i d is equal to zero when A1 =A2, maintaining
the thyristor in conduction state even when Vs < 0. When an inductive active load
is connected to the rectifier, as illustrated in Fig. 15-b, the thyristor will be turned on if the
firing pulse is applied to the gate when Vs > Ed . Again, the thyristor will remain in the on-
state until A1 = A2. When the thyristor is turned off, the load voltage will be Vd = Ed .
Fig. 15 single thyristor rectifier with: (a) R-L load (b) active load
96
is = iT1 – iT4 (9)
Fig. 16 Single phase bridge rectifier (a) fully controlled (b) half controlled
and its waveform is shown in Fig. 17. Figure 18 presents the behavior of the fully controlled
rectifier with resistive–inductive load (with L ). The high load inductance generates a
perfectly filtered current and the rectifier behaves like a current source. With continuous
load current, thyristors T1 and T2 remain in the on-state beyond the positive half-wave of
the source voltage Vs . For this reason, the load voltage Vd can have a negative instantaneous
value. The firing of thyristors T3 and T4 has two effects:
(i) They turn-off thyristors T1 and T2
(ii) After the commutation, they conduct the load current.
This is the main reason why this type of converters are called “naturally commutated” or
“line commutated” rectifiers. The supply current iS has the square waveform, as shown in Fig.
18, for continuous conduction. In this case, the average load voltage is given by
Vdiα = ∫ (10)
97
Fig. 17 Waveforms of a fully controlled bridge rectifier with R-load
98
Fig. 18 Waveforms of a fully controlled bridge rectifier with R-L load (L )
99
Single-phase inversion circuits
The main objective of static power converters is to produce an Ac output waveform from a dc
power supply. These are the types of waveforms required in adjustable speed drives (ASDs),
uninterruptible power supplies (UPSs), static var compensators, active filters, flexible ac
transmission systems (FACTSs), and voltage compensators, which are only a few applications.
For sinusoidal Ac outputs, the magnitude, frequency, and phase should be controllable.
According to the type of ac output waveform, these topologies can be considered as voltage-
source inverters (VSIs), where the independently controlled ac output is a voltage waveform.
These structures are the most widely used because they naturally behave as voltage sources
as required by many industrial applications, such as ASDs, which are the most popular
application of inverters (Fig. 19-a). Similarly, these topologies can be found as current-source
inverters (CSIs), where the independently controlled Ac output is a current waveform. These
structures are still widely used in medium-voltage industrial applications, where high-quality
voltage waveforms are required. Static power converters, specifically inverters, are
constructed from power switches and the Ac output waveforms are therefore made up of
discrete values. This leads to the generation of waveforms that feature fast transitions rather
than smooth ones. For instance, the Ac output voltage produced by the VSI of a three-level
ASD is a, Pulse Width Modulation (PWM) type of waveform (Fig. 19). Although this waveform is
not sinusoidal as expected (Fig. 19), its fundamental component behaves as such. This
behavior should be ensured by a modulating technique that controls the amount of time and
the sequence used to switch the power valves on and off.
Fig. 19 A three level adjustable speed drive scheme and associated waveforms (a) power
conversion topology (b) ideal input and output waveforms (c) actual input and output
100
waveforms
The modulating techniques most used are the carrier-based technique (e.g. sinusoidal pulse
width modulation, SPWM), the space-vector (SV) technique, and the selective-harmonic-
elimination (SHE) technique.
The discrete shape of the ac output waveforms generated by these topologies imposes basic
restrictions on the applications of inverters. The VSI generates an Ac output voltage waveform
composed of discrete values (high dv/dt ); therefore, the load should be inductive at the
harmonic frequencies in order to produce a smooth current waveform. A capacitive load in
the VSIs will generate large current spikes. If this is the case, an inductive filter between the
VSI Ac side and the load should be used. On the other hand, the CSI generates an Ac output
current waveform composed of discrete values (high di/dt ); therefore, the load should be
capacitive at the harmonic frequencies in order to produce a smooth voltage waveform. An
inductive load in CSIs will generate large voltage spikes. If this is the case, a capacitive filter
between the CSI Ac side and the load should be used. A three-level voltage waveform is not
recommended for medium-voltage ASDs due to the high dv/dt that would apply to the motor
terminals. Several negative side effects of this approach have been reported (bearing and
isolation problems). As alternatives, to improve the ac output waveforms
in VSIs are the multistage topologies (multilevel and multi-cell). The basic principle is to
construct the required Ac output waveform from various voltage levels, which achieves
medium-voltage waveforms at reduced dv/dt . Although these topologies are well developed
in ASDs, they are also suitable for static var compensators, active filters, and voltage
compensators. Specialized modulating techniques have been developed to switch the higher
number of power valves involved in these topologies. Among others, the carrier-based (SPWM)
and SV-based techniques have been naturally extended to these applications.
In many applications, it is required to take energy from the ac side of the inverter and send it
back into the dc side. For instance, whenever ASDs need to either brake or slow down the
motor speed, the kinetic energy is sent into the voltage dc link (Fig. 19-a). This is
known as the regenerative operating mode and, in contrast to the motoring mode, the dc link
current direction is reversed due to the fact that the dc link voltage is fixed. If a capacitor is
used to maintain the dc link voltage (as in standard ASDs) the energy must either be
dissipated or fed back into the distribution system, otherwise, the dc link voltage gradually
increases. The first approach requires the dc link capacitor be connected in parallel with a
resistor, which must be properly switched only when the energy flows from the motor into the
dc link. A better alternative is to feed back such energy into the distribution system.
However, this alternative requires a reversible-current topology connected between the
101
distribution system and the dc link capacitor. A modern approach to such a requirement is to
use the active front-end rectifier technologies, where the regeneration mode is a natural
operating mode of the system.
Half-bridge VSI
Figure 20 shows the power topology of a half-bridge VSI, where two large capacitors are
required to provide a neutral point N, such that each capacitor maintains a constant voltage
Vi /2. Because the current harmonics injected by the operation of the inverter are low-order
harmonics, a set of large capacitors (C+ and C−) is required. It is clear that both switches S+
and S− cannot be on simultaneously because a short circuit across the dc link voltage source Vi
would be produced. There are two defined (states 1 and 2) and one undefined (state 3) switch
state as shown in Table 1. In order to avoid the short circuit across the dc bus and the
undefined Ac output-voltage condition, the modulating technique should always ensure that
at any instant either the top or the bottom switch of the inverter leg is on.
102
Figure 21 shows the ideal waveforms associated with the half-bridge inverter shown in Fig. 20.
The states for the switches S+ and S− are defined by the modulating technique, which in this
case is a carrier-based PWM.
̇ ̇ (13)
for ma ≤ 1, which is called the linear region of the modulating technique (higher values of ma
leads to over-modulation)
103
Fig. 21 The half bridge VSI waveform for the SPWM (ma =0.8, mf =9) (a) carrier and
modulating signals (b) S+ state (c) S- state (d) Ac output (e) Ac output spectrum (f) Ac output
current (g) dc current (h) dc current spectrum (i) switch S+ current (j) D + current
̇ ̇ (14)
Fig. 22 shows output voltage of the half bridge VSI, for square wave modulation technique.
104
Fig. 22 The half bridge VSI, for square wave modulation technique
Full-bridge VSI
Figure 23 shows the power topology of a full-bridge VSI. This inverter is similar to the half-
bridge inverter; however, a second leg provides the neutral point to the load. As expected,
both switches S1+ and S1− (or S2+ and S2−) cannot be on simultaneously because a short
circuit across the dc link voltage source Vi would be produced. There are four defined (states
1, 2, 3, and 4) and one undefined (state 5) switch state as shown in Table 2. The undefined
condition should be avoided so as to be always capable of defining the Ac output voltage
always. In order to avoid the short circuit across the dc bus and the undefined Ac output
voltage condition, the modulating technique should ensure that either the top or the bottom
switch of each leg is on at any instant. It can be observed that the Ac output voltage can take
values up to the dc link value Vi, which is twice that obtained with half-bridge VSI topologies.
Several modulating techniques have been developed that are applicable to full-bridge VSIs.
Among them are the PWM (bipolar and unipolar) techniques.
Bipolar PWM Technique
States 1 and 2 (Table 2) are used to generate the Ac output voltage in this approach. Thus,
the Ac output voltage waveform features only two values, which are Vi and −Vi. To generate
the states, a carrier-based technique can be used as in half-bridge configurations (Fig. 20),
where only one sinusoidal modulating signal has been used. It should be noted that the on-
state in switch S+ in the half-bridge corresponds to both switches S1+ and S2− being in the on-
state in the full-bridge configuration. Similarly, S− in the on-state in the half-bridge
corresponds to both switches S1− and S2+ being in the on-state in the full-bridge configuration.
This is called bipolar carrier-based SPWM.
105
Fig. 23 Single phase full bridge VSI
The Ac output voltage waveform in a full-bridge VSI is basically a sinusoidal waveform that
features a fundamental component of amplitude ˆvo1 that satisfies the expression
̇ ̇ (16)
In the linear region of the modulating technique (m a ≤ 1), which is twice that obtained in the
half-bridge VSI. Identical conclusions can be drawn for the frequencies and the amplitudes of
the harmonics in the Ac output voltage and dc link current, and for operations at smaller and
larger values of odd mf (including the over-modulation region (ma > 1)), than in half-bridge
VSIs, but considering that the maximum ac output voltage is the dc link voltage Vi . Thus, in
the over-modulation region the fundamental component of amplitude ̇ satisfies the
expression
̇ ̇ (17)
106
Chapter 5
Control System Modeling and Representation
Objectives
Presentation of the main mathematics required for modeling, control systems.
Providing methods and tools for proper modeling and representation of physical
systems.
Introduction
The Laplace transform is an operational method that can be used for solving linear
differential equations. By use of Laplace transforms, we can convert many common functions,
such as sinusoidal functions, damped sinusoidal functions, and exponential functions, into
algebraic functions of complex variables. Operations such as differentiation and integration
can be replaced by algebraic operations in the complex plane. Therefore, a linear differential
equation can be transformed into an algebraic equation in a complex variable s. If the
algebraic equation in s is solved for the dependent variable, then the solution of the
differential equation (the inverse Laplace transform of the dependent variable) may be found
by use of a Laplace transform table or by use of the partial-fraction expansion technique. An
advantage of the Laplace transform method is that it allows the use of graphical technique for
predicting the system performance without actually solving system differential equations.
Another advantage of the Laplace transform method is that, when we solve the differential
equation, both the transient component and steady-state components of the solution can be
obtained simultaneously. Laplace transform is followed by a discussion of aspects of modeling
control system components and an introduction to the concept of transfer functions.
Representation of control systems using block diagrams is covered to conclude this chapter.
The Laplace transform is one of the most indispensable tools at the disposal of the control
systems engineering. The fundamental definition of the Laplace transform X(s) of a time-
varying function x(t) is given by the basic relationship
(1)
∫
107
Note that the operation on the right-hand side of Equation (1) involves the multiplication of
the variable function x(t), which is assumed to be defined from t = 0 to infinity by the factor
e-st. The new variable (s) is called the Laplace operator. The resulting expression is clearly a
function of s only, as the integration is between zero and infinity. A commonly used notation
to express the Laplace transform operation is
{ } (2)
The Laplace transform of some common functions are obtained in the following examples.
Table 1 lists the Laplace transform of functions encountered in control systems engineering.
TABLE 1
SOME LAPLACE TRANSFORM PAIRS
X(t) X(s)
Unit impulse δ(t) 1
Unit step u(t)
sin ωt
cos ωt
sin ωt
cos ωt
Example 1
Find the Laplace transform of the function
x(t) = eλt t≥0
Solution
Using Eq. (1), we can write
∫
Thus
108
We can thus conclude that
(3)
{ }
The result of Example 1 is useful in many ways. First, the Laplace transform of the
exponential function is important and should be memorized. Second, Eq. (3) can be used to
derive the Laplace transform of other functions. To start, note that a step function u(t) is
defined as
u(t) = 1 t≥0
u(t) = 0 t≤0
Note that
u(t) = e0t (4)
∫[
∫ ∫
Thus we conclude that
{ } { } { } (6)
We can combine (6) with the result of Example 1 to derive the Laplace transform of other
functions.
Example 2
Find the Laplace transform of
109
Solution
First recall Euler’s identity,
Thus
where, √ .
We can thus write
[ ]
Clearly we have
( )
( ) ( )
Thus,
(7)
{ }
If α = 0, we obtain
110
{ } (8)
Application of the Laplace transform to both sides, assuming zero initial conditions, gives us
Suppose now that the input function u(t) is a unit step; thus
Finding the function x(t) whose transform is as given above, is symbolized by the inverse
transform operator -1; Thus
{ } { }
{ } ∫ (10)
(11)
111
The functions X1(s), X2(s), … , Xn(s) can then be looked up in a table of Laplace transform pairs
and hence we can obtain the corresponding inverse x1(t), x2(t), … , xn(t). The final result is
then
(12)
where, A(s) and B(s) are polynomials in s. In the expansion of F(s) = B(s)/A(s) into a partial-
fraction form, it is important that the highest power of s in A(s) be greater than the highest
power of s in B(s). If such is not the case, the numerator B(s) must be divided by the
denominator A(s) in order to produce a polynomial in s plus a remainder. (This remainder is a
ratio of polynomials in s whose numerator is of lower degree than the denominator.) If F(s) is
broken up into components,
and if the inverse Laplace transforms of F1(s), F2(s), …, Fn(s) are readily available, then
[ [ [ [
[ (13)
where, f1(t), f2(t), …, fn(t) are the inverse Laplace transform of F1(s), F2(s), …, Fn(s),
respectively. The inverse Laplace transform of F(x) thus obtained is unique except possibly at
points where the time function is discontinuous. Whenever the time function f(t) and its
Laplace transform F(s) have a one-to-one correspondence. The advantage of the partial-
fraction expansion approach is that the individual terms of F(s), resulting from the expansion
into partial fraction form, are very simple functions of s; consequently, it is not necessary to
refer to a Laplace transform table if we memorize several simple Laplace transform pairs. It
should be noted, however, that in applying the partial fraction expansion technique in the
search for the inverse Laplace transform of F(s) = B(s)/A(s) the roots of the denominator
polynomial A(s) must be obtained in advance. That is, this method does not apply until the
denominator polynomial has been factored.
112
for m < n
where, p1, p2, …,pn and z1, z2, …, zm are either real or complex quantities, but for each
complex pi or zj there will occur the complex conjugate of pi or zj, respectively.
If F(s) involves distinct poles only, then it can be expanded into a sum of simple partial
fractions as follows:
(14)
The coefficient ak is called the residue at the pole s =-pk. The value of ak can be found by
multiplying both sides of equation (2-14) by (s + pk) and letting s = -pk, which gives
[ ]
We see that all the expanded terms drop out with the exception of ak. Thus the residue ak is
found from
[ ]
(15)
Note that, since f(t) is a real function of time, if p1 and p2 are complex conjugates, then the
residues a1 and a2 are also complex conjugates. Only one of the conjugates a1 or a2 needs to
be evaluated because the other is known automatically.
[ ]
f(t) obtained as
[ for t ≥ 0
113
Example 3
Find the inverse Laplace transform of
Solution
The partial fraction expansion of F(s) is
[ ] [ ]
[ ] [ ]
Thus
[ [ ] [ ]
for t ≥ 0
Example 4
Find the inverse Laplace Transform of
Solution
We write
Thus
Put
s = -2; A1 = 1/27
Put
s = -5; A2 = -1/18
Put
114
s = -11; A3 = 1/54
As a result,
( )
Example 5
Obtain the inverse Laplace transform of
Solution
Here, since the degree of the numerator polynomial is higher than that of the denominator
polynomial, we must divide the numerator by the denominator
s +2
s + 3s + 2 s3 + 5s2 + 9s + 7
2
s3 + 3s2 + 2s
0 + 2s2 + 7s + 7
2s2 + 6s + 4
0 + s +3
Note that the Laplace transform of the unit impulse function δ(t) is 1, and that the Laplace
transform of dδ/dt is s.The third term on the right hand side of the last equation is F(s) in
Example 2-3. So the inverse Laplace transform of G(s) is given as
for t ≥ 0
Example 6
Find the inverse Laplace transform of
Solution
115
Notice that the denominator polynomial can be factored as
If the function F(s) involves a pair of complex-conjugate poles, it is convenient not to expand
F(s) into partial fractions but to expand it into the sum of a damped sine and a damped cosine
function. Noting that s2 + 2s + 5 = (s + 1)2 + 22 and referring to the Laplace transform of e-
αt
sinωt and e-αtcosωt, rewritten thus,
{ }
{ }
The given F(s) can be written as a sum of a damped sine and a damped cosine functions
It follows that
{ }
[ ] [ ]
Instead of discussing the general case, we shall use an example to show how to obtain
the partial fraction expansion of F(s). Consider the following F(s).
116
By multiplying both sides of the last equation by (s+1)3, we have
(16)
[ ]
[ ]
(17)
[ ]
[ ]
From the preceding analysis it can be seen that the values of b3, b2 and b1 are
found systematically as follows:
[ ]
{ [ ]} [ ]
{ [ ]} [ ]
We thus obtain
117
[ ] [ ] [ ]
for t ≥ 0
Example 7
Find the Laplace transform of the following function
Solution
We write
Thus
Put
s = 0; A2 = 1/3
Put
s = -3; A3 = 1/9
Thus
A1 = - 1/9
As a result
( )
Example 8
Find the Laplace transform of the following function
Solution
118
We can write
Thus
Put
s = 0; A1 = 1/9
Put
s = -3; A3 = -1/3
Thus
A2 = - 1/9
As a result,
[ ]
The Laplace transform can be used for solving linear time-invariant differential equations
with known initial conditions. In solving differential equations two steps are involved.
Step 1. By taking the Laplace transform of each term in the given differential
equation. Convert the differential equation into an algebraic equation in s. Obtain
the expression for the Laplace transform of the dependant variable by rearranging
the algebraic equation.
Step 2. The time solution of the differential equation is obtained by finding the
inverse Laplace transform of the dependant variable. In the following discussion, two
examples are used to demonstrate the solution of linear, time-invariant, differential
equation by the Laplace transform method.
Example 9
Find the solution x(t) of the differential equation
̈ ̇ =0 ̇
119
[
we obtain
[ ̇
[ ̈ ̇
[ ̇ [
By substituting the given initial conditions into this last equation, we obtain
[ [
Or
[ ] [ ]
, for t ≥ 0
Example 10
Find the solution x(t) of the differential equation
̈ ̇ ̇
Solution
Noting that [ , , and ̇ ,
The Laplace transform of the differential equation becomes
120
Hence the inverse Laplace transform becomes
[ ] [ ] [ ]
for t ≥ 0
Control system engineers use physical laws that describe the interaction between variables of
interest in the system under consideration. We use the simple laws of the system elements. In
this lecture we discuss several laws of simple elements.
Passive Elements
Passive elements in electric circuits are the resistance R, the inductance L, and the
capacitance C. The relation between the voltage across v(t) and current through i(t)
depends on the element.
For a resistance element we have
121
Kirchhoff’s laws are also useful (described in a later chapter).
Transfer Functions
A transfer function is defined as the ratio of the Laplace transform of the output variable
in a linear time-invariant dynamic system to the Laplace transform of the input variable with
zero initial conditions. A number of examples is given below, together with the derivations
associated with simple dynamic systems.
Figure 1 shows a diagram of separately excited dc generator. The input voltage vi(t)
produces a current if(t) in the field circuit.
Assuming that the field circuit resistance is Rf and the inductance is Lf, we can write
Assuming that the magnetization characteristic of the machine is linear in the region of
interest, and that the generator is driven at constant speed, we can write the output voltage
as
As a result, the transfer function between input and output voltages is given by
Amplidyne
122
Figure 2. The amplidyne.
Assume that the control winding’s resistance and inductance are denoted as Rc and Lc. Let the
control (input) voltage is vc(t) and the current in the control winding is ic(t). We have
The electromotive force (emf) developed in the quadrature axis winding is proportional to the
field current.
Thus
The quadrature-axis current sets up a flux, which in turn produces the output voltage vd (t)
according to
The Laplace domain equation for the above time domain equations are:
As a result, we have
( )
which is the transfer function of the amplidyne.
In a field controlled dc motor (Figure 3), the voltage input is fed to the field, which can
be modeled as a resistance Rf in series with the inductance Lf. A field current if(t) is
established in accordance with
123
The field current sets up a flux which together with the rotational speed of the motor
develops a back emf em(t).
The parameter Km is a constant that depends on the motor’s design particulars. The motor’s
developed electrical power is given in terms of the armature current Ia and the back emf as
Note that the armature current is assumed constant. We thus can write
Assuming that the electrical power Pe(t) suffers no losses in being transmitted as mechanical
power Pm(t), we can write
We know that the mechanical power is the product of torque and angular velocity, thus
It is thus clear that under the foregoing assumptions, we can write the developed torque
as
where,
As a result, we conclude that the motor’s torque is proportional to the field current. The
torque equation; expressed in the Laplace transform is
The relation between the field current and the input voltage is
As a result we conclude that the transfer function between the input voltage and output
torque can be written as
If the input voltage to the dc motor is fed to the armature circuit, while field current is
maintained constant, we say that the motor is armature controlled.
124
Figure 4. Schematic diagram of an armature-controlled dc motor.
In this equation Ra and La are the resistance and the inductance of the armature circuit,
respectively.
Thus
The armature current is denoted by ia(t), while the motor’s back emf is denoted by em(t). The
input voltage is denoted by vi(t). Since the field current is fixed at If, we write:
The torque developed by the motor under assumptions similar to those stated for the field
controlled dc motor can be obtained from
Thus
As a result, we write
Thus the back emf and torque are proportional to the motor’s velocity and armature current,
respectively. From the previous discussions we can write the previous equations transformed
into Laplace domain.
125
Thus the Laplace transform of the developed torque is given by:
This equation is not a single-input output relationship and thus a straightforward transfer
function cannot be obtained for T(s) in terms of the input voltage. It is possible, however, to
obtain a transfer function between ωm(s) and Vi(s) as indicated below. Assume that the motor
is driving a load such that no opposing torque Ti exists. We can thus write
( )
where, Jeq is the moment of inertia of the rotating parts, and Beq is a term denoting friction
coefficient.
( )
or,
[( ) ]
As a result, a transfer function between ωm(s) as output and Vi(s) as input can be written as:
( )
[ ]
[ ]
( )
( )
Two-Phase Servomotor
126
In a two-phase servomotor (Figure 5a) the torque decreases linearly with the speed increase
for a constant control winding voltage.
We know that the speed ω is equal to the time derivative of the angle θ. Thus the last
equation can be rewritten as:
̈ ̇
where J is the moment of inertia of the motor and load referred to the motor shaft, and B is
the friction coefficient.
̈ ̇
Noting that the control voltage ec(t) is the input and the displacement angle θ is the output,
we see that the transfer function of the system is given by:
⁄
[ ⁄
127
⁄
{[ ⁄ }
where, Km= Kc / (B + Kn) = motor gain constant, and Tm = J / (B + Kn) = motor time constant
A diagram showing a mass is given in Figure 6a. The elastance, or stiffness, K of a spring
provides a reaction force that is proportional to the deformation Δx of the spring.
FK = K Δx
Note that in Figure 6b,
Δx =x1 –x2
Viscous friction or damping B involves energy absorption. The damping force is proportional to
the difference of the velocities of the two bodies (Figure 6c).
128
acceleration corresponds to the translational acceleration a. Torque equations for rotational
systems are parallel to force equations in transitional systems. For a body with a moment of
inertia J, an applied torque T produces an angular acceleration d2θ/dt2. The reaction torque
TJ is in opposition to T, with a torque equation given by
Gear Trains
A normal practice in coupling motors to loads is to employ a gear train to transmit the
driving torque to the load. An analysis of such a case is important to evaluate the moment of
inertia and damping of the load relative to the motor. Figure 8 shows a driving motor that
supplies input torque Tm at an angular velocity ωm to a load through a gear train having a gear
ratio of Nm/Nl. The load torque is denoted by Tl and its angular velocity is denoted by ωl.
The inertia and viscous friction of the motor are denoted by Jm and Bm. In a similar way Jl and
Bl are moment of inertia and viscous friction of the load. The torque balance equation on the
motor side is given by:
̈ ̇
Thus the torque provided by the motor is equal to the sum of motor inertial and friction
torques and the torque Tml transmitted through the gears. Assuming no power loss in the
129
gears, we can write the power equation:
This is a restatement of the fact that power is the product of torque and angular velocity. We
also know that:
Now the load side a driving torque balance equation can be written as
̃ ̈ ̇
In terms of the motor’s angular velocity ̇ and angular acceleration we thus have
̃ ( ̈ ̇ )
Or
̃ ( ̈ ̇ )( )
As a result, we can assert the motor’s driving torque is given by
( ( ) ) ̈ ( ( ) ) ̇
This result shows that an equivalent moment of inertia Jeq and an equivalent viscous friction
Beq are experienced by the motor:
( )
( )
The load torque is seen by the motor as
130
̃
A transfer function has been stated as the ration of the Laplace transform of the output
variable in a linear time invariant dynamic system to the Laplace transform of the input with
zero initial condition. Here is a number of applications.
RC Circuit
( )
131
Consider the RC Differentiating circuit of Figure 10.
( )
132
⁄
Note the similarity of this transfer function and that of the RC integrating circuit. Consider
the Spring-Dashpot system of Figure 11, which has B and K interchanged, as shown in Figure
12. The force balance equation is given by:
̇ ̇
133
The above function is similar to that of the differentiating RC circuit.
We can write:
Thus
The transfer function obtained, relates the transferred output velocity to the transform of the
input torque, in the presence of the moment of inertia J and viscous friction B.
The system consists of a spring-mass-dashpot combination. We can write the force balance
equation as
134
̈ ̇
( )
135
Figure 16. Lead-lag RC circuit
⁄
⁄
Let
Thus
Also
Let
Thus
136
Let ;
thus
Where
Problems
f (t ) 0 for t 0
a)
te 3t , for t 0
f (t ) 0 for t 0
b)
t 2 e at , for t 0
f (t ) 0 for t 0
c)
3 sin(5t 45), for t 0
2) Find the inverse Laplace Transform of F(s), where
137
1
F ( s)
s( s 2 2s 2)
3) Obtain the inverse Laplace transform of
5( s 2)
F ( s) 2
s ( s 1)(s 3)
4) Find the inverse Laplace Transform of
s 4 2s 3 3s 2 4s 5
F ( s)
s( s 1)
5) Solve the differential equation:
x.. 2x. 10x t 2 , x(0) 0, x. (0) 0
138
Chapter 6
Block Diagrams and Analysis of the Responses
Objectives
Basics of block diagrams.
Representation of transfer functions.
Transfer functions presentation of some practical systems.
Definition
1. Functional block
This is a symbol representing the transfer between the input U(s) to an element and the
output X(s) of the element. The block contains the transfer function G(s), as shown in Figure
1.
The arrow directed into the block represents the input U(s), while that directed out of the
block represents the output X(s). The block shown in the figure represents the algebraic
relationship
X(s) = G(s) U(s)
2. Summing point
This is a symbol denoted by a circle, the output of which is the algebraic sum of the
signals entering into it. A minus sign close to an input signal arrow denotes that this signal is
reversed sign in the output expression. Figure 2 shows the relationship
E(s) = R(s) – C(s)
139
Figure 2. Summing point.
A fundamental block diagram configuration is the single-loop feedback system shown in Figure
4a. The output variable C(s) is modified by the feedback element with transfer function H(s)
to produce the signal B(s):
B(s) = C(s) H(s)
The signal B(s) is compared to a reference signal R(s) to produce the error signal E(s).
E(s) = R(s) – B(s)
The error signal actuates the plant with transfer function G(s) to produce the output C(s):
C(s) = G(s) E(s)
140
(C(s)/G(s)) = R(s) – C(s) H(s)
( )
To obtain the overall relationship between the outputs and inputs of complex systems, we
often have to eliminate variables in the system representation. We consider here the transfer
functions of cascaded elements as shown in Figure 5a. We write
X2 = G 1 X1
X3 = G2 X 2 = G1 G 2 X1
141
Figure 5. Cascaded system and the steps in its reduction.
Table 1.
Block diagram reduction equivalents
Original Configuration Alternate Configuration
(a)
(b)
142
(c)
(d)
(e)
(f)
(g)
143
(h)
Example 1.
Simplify the block diagram of Figure 5.6.
Solution
First, move the branching point b to a as shown in Figure 7(a).
Figure 7(a)
Now we see that G1 and G2 are in parallel, and we reduce the figure to that of Figure
7(b).
144
Figure 7(b)
The feedback loop with a forward gain of 1 and feedback element G2H can be reduced as
shown in Figure 7(c).
Figure 7(c).
Figure 7(d).
Example 2
Use block diagram reduction techniques to obtain the ratio C/R for the system shown in the
block diagram of Figure 8.
145
Figure 9. Steps in the foregoing Figure 5.8.
( )
This is represented in block diagram form in the first block of Figure 10.
Assume now that the motor is driving a load torque Tl through a gear train of speed ration
Nm/Nl . The motor’s inertia and viscous friction are Jm and Bm , and the load’s inertia and
viscous friction are Jl and Bl. We know that
146
̈ ̇ ̃
In the s domain this is written as
( ) ̃
This equation is realized in block diagram form by the right-hand block if Figure 5.10.
( ) ̃
( ) ̃
A block diagram can be constructed as shown in Figure 11a. Note the presence of a feedback
path to account for the effect of the motor’s velocity on armature current. By moving the
load torque summing junction to the left- hand side as shown in Figure 11b, we can see that
the motor is in actual fact subject to two inputs: Vi(s) and ̃ .
(a)
(b)
Figure 11. Block Diagram of an Armature-Controlled DC Motor.
147
Figure 12. Block Diagram of DC Generator.
The commonly used test input signals are those of step functions, acceleration functions,
sinusoidal functions, and the like. With these test signals, mathematical and experimental
analyses of control systems can be carried out easily since the signals are very simple
functions of time. Which of these typical input signals to use for analyzing system
characteristics? The answer is determined by the form of the input that the system will be
subjected to most frequently under normal operation. If the inputs to a control system are
gradually changing functions of time, then a ramp function of time may be a good test signal.
Similarly, if a system is subjected to sudden disturbances, a step function of time may be a
good test signal.
For a system subjected to shock inputs, an impulse function may be best. Once a control
system is designed on the basis of test signals, the performance of the system in response to
actual inputs is generally satisfactory. The use of such test signals enables one to compare the
performance of all systems on the same basis.
The time response of a control system consists of two parts: the transient response and
the steady state response. By transient response, we mean that which goes from the initial
state to the final state. By steady-state response, we mean the manner in which the system
output behaves as t approaches infinity. Thus, the system response can be written as
148
where, ctr(t) is the transient response, and css(t) is the steady-state response.
The most important characteristic of the dynamic behavior of a control system is absolute
stability, i.e. whether the system is stable or unstable. A control system is stable if, in the
absence of any disturbance or input, the output stays in the same state. A linear control
system is stable if the output eventually comes back to its equilibrium state when the system
is subjected to an initial condition. A linear control system is critically stable if it oscillations
of the output continue forever. It is unstable if the output divers without bound from its
equilibrium state when the system is subjected to an initial condition. Important system
behavior includes relative stability and steady-state error.
Since a physical control system involves energy storage, the output cannot follow the
input immediately but exhibits a transient response before a steady-state can be reached.
The transient response often exhibits damped oscillations before reaching steady state. If the
output of a system at steady state does not exactly agree with the input, the system is said to
have steady-state error. This error is indicative of the accuracy of the system. In analyzing a
control system, we must examine transient-response behavior and steady-state behavior.
First-Order Systems
Consider the block diagram of a first-order system shown in Figure 13a. Physically this block
diagram represents an RC circuit, or a thermal system. A simplified block-diagram is shown in
Figure 13b. The input-output relationship is given by:
We shall analyze the system responses to such inputs as the unit step, unit ramp, and unit
impulse functions. The initial conditions are assumed to be zero. The following analysis is
correct, whatever the physical system is, as long as the transfer function is the same.
149
Figure 13. Block diagram of a first-order system.
Since the Laplace transform of the unit-step function is 1/s, substituting R(s) = 1/s into the
above equation, we obtain
for t ≥ 0
The last equation states that initially the output c(t) is zero and finally it becomes unity. One
important characteristic of such an exponential response curve c(t) is that at t = T the value
of c(t) is 0.632, or the response has reached 63.2% of its total change. This may be easily seen
by substituting t =T in c(t).
It is to be noted that the smaller the time constant T, the faster the system response. Another
important characteristic of the exponential response curve is that the slope of the tangent
line at t = 0 is 1/T, since
The output would reach the final value at t=T if it maintained its initial speed of response.
From the preceding exponential equation we see that the slope of the response curve c(t)
decreases monotonically from 1/T at t = 0 to zero at t = . The exponential response curve is
shown in Figure 14.
150
Figure 14. Exponential response curve,
In one time constant, the exponential response curve has gone from 0 to 63.2% of the final
value. In two time constants, the response reaches 86.5% of the final value. At t = 3T, 4T, 5T,
the response reaches 95%, 98.2% and 99.3%, respectively, of the final value. Thus, for t ≥ 4T,
the response remains within 2% of the final value. The exponential equation states that the
steady-state is reached only at t = ∞.
In practice, a reasonable estimate of the response time is the length of time the
response curve needs to reach and stay within 2% less than the final value, or four time
constants.
Since the Laplace transform of the unit-ramp function is 1/s2, we obtain the output of the
system of Figure 13 as:
, for t ≥ 0.
151
( )
As t approaches infinity, e-t/T approaches zero, and thus the error signal e(t) approaches T, or
The unit-ramp input and the unit output are shown in Figure 15. The error in following the
unit-ramp input is equal to T for sufficiently large t. The smaller the time constant T, the
smaller the steady-state error in following the ramp input.
For the unit-impulse input, R(s) = 1 and the output of the system of Figure 13 can be
obtained as:
for t ≥ 0
The response is given by the following equation and is shown in Figure 16.
152
Figure 16. Unit impulse response of the first-order system.
In the above analysis, it has been shown that for the unit-ramp input the output c(t) is
, for t ≥ 0.
For the unit-step input, which is the derivative of unit-ramp input, the output c(t) is
for t ≥ 0
Finally, for the unit-impulse input, which is the derivative of unit step input, the output c(t) is
for t ≥ 0
Comparing the system response to these three inputs clearly indicates that the response
to the derivative of an input signal can be obtained by differentiating the response of the
system to the original signal. It can be also seen that the response to the integral of the
original signal can be obtained by integrating the response of the system to the original and
by determining the integration constant from the zero output initial condition. Systems
153
References and Recommended Readings
Main References
[1] Ogata K, Yang Y. Modern control engineering. India: Prentice hall; 2002.
[2] Luyben WL. Process modeling, simulation and control for chemical engineers. McGraw-
Hill Higher Education; 1989 Aug 1.
[3] Floyd TL, Buchla DM. Electric circuits fundamentals. Pearson/Prentice Hall; 2004.
[4] Charles K. Alexander & Matthew N. O. Sadiku. Fundamentals of Electric Circuits, 4th
edition, McGraw-Hill, 2009.
[5] Muhammad H. Rashid. POWER ELECTRONICS HANDBOOK DEVICES, CIRCUITS, AND
APPLICATIONS, Third Edition, Elsevier Inc., 2011.
Recommended Readings
[1] Golnaraghi F, Kuo BC. Automatic control systems. Complex Variables. 2010.
[2] Lewis PH, Yang C. Basic control systems engineering. 1997
[3] Nilsson JW. Electric circuits. Pearson Education India; 2008.
[6] Nilsson JW, Riedel S. Electric circuits. Prentice Hall Press; 2010 Jan 13.
[4] Jackson HW, Temple D, Kelly B. Introduction to electric circuits. Englewood Cliffs:
Prentice-Hall; 1986.
[5] Ned Mohan. Power electronics : a first course, John Wiley & Sons, Inc., 2012.
[6] Richard C. Dorf, James A. Svoboda . Introduction to Electric Circuits, EIGHTH EDITION,
John Wiley & Sons, Inc., 2010.
154