ELECTRONICS
Explained Simply and Easily
A Quick Guide to Electronic Fundamentals,
Components, Circuits and Applications
Brought to you by
VPACHKAWADE Research Center
www.vpachkawade.com
[email protected]
No part of this book may be reproduced, distributed, or transmitted in any
form or by any means, including photocopying, recording, or other
electronic or mechanical methods, without the prior written permission of
the publisher, except in the case of brief quotations embodied in critical
reviews and certain other noncommercial uses permitted by copyright law.
Disclaimer of Liability:
The author and publisher have made every effort to ensure the accuracy of
the information in this book but do not guarantee its completeness or
suitability for any purpose. This book is provided on an "as-is" basis, and
the author and publisher shall not be held liable for any damages or losses
arising from its use.
Trademark Acknowledgment:
All trademarks and brand names mentioned in this book are the property of
their respective owners.
For permissions, inquiries, or feedback, please contact:
[email protected]
Page intentionally left blank
Table of Contents
Chapter 1 – Introduction
Chapter 2 – Basic electronic components
Chapter 3 – Electronic fundamental
Chapter 4 – Semiconductor
Chapter 5 – Semiconductor devices
Chapter 6 – Diode and Transistor circuits
Chapter 7 – Principles of electronic measurements
Chapter 8 – Additional topics
What will you learn from this book?
Learners can expect to achieve following learning objectives or
outcomes after reading this book.
Learn about electronic components, their symbols, and
functions within circuits
Gain proficiency in analyzing electronic components and
circuits using fundamental laws and techniques
Learn details on KVL, KCL, diode, diode circuits, transistor
and transistor circuits, biasing methods, etc.
Learn basics and advances of electronic components,
semiconductor devices (diode, bjt, mosfet, op-amp, etc)
Develop skills in designing and prototyping electronic
circuits on breadboards and troubleshooting circuit designs
Learn skills to test and make measurements in electronics
design
Learn how to use software (CAD) to design, simulate and
analyse circuits
Learn practical electronics - do it yourself practice
Are there requirements/prerequisites for
reading this book?
No specific requirements or prior experience necessary - this
beginner-friendly book on Electronics welcomes all
interested learners.
Who is this book for?
Below is the description of the intended learners for our book who will
find our book content valuable.
This course is intended for beginners and enthusiasts who
want to learn about Electronics.
Undergraduate/graduate students in engineering, science and
technology in all disciplines
Electronics hobbyists
Engineers and technicians
Anyone curious about how electronics work and how to use
and apply it to develop practical applications.
Whether you have a background in electronics or are
completely new to the subject, this course will provide a
solid foundation and practical knowledge to start working
with electronics and undertake exciting projects.
Professionals willing to revise the concepts in electronic
components, semiconductor devices and circuits analysis
Book Description
Are you eager to understand the world of electronics? Whether you're an
aspiring engineer, a tech enthusiast, or someone simply curious about how
the gadgets you use every day work, our "Electronics - A Quick Guide to
Electronic Fundamentals, Circuits and Applications" is your ticket to
understanding the fundamental principles that power the modern world.
Book Highlights:
Foundations of Electronics: Dive into the very core of electronics by
exploring concepts like voltage, current, resistance, capacitance and power.
Understand Ohm's law and how it governs the behavior of electronic
components.
Electronic Components: Discover the key building blocks of electronic
circuits, including resistors, capacitors, inductors, diodes, and transistors.
Learn how to identify, use, and connect these components effectively.
Circuit Analysis: Master the art of analyzing simple electronic circuits.
You'll develop the skills to calculate voltages, currents, and power across
different components using various circuit analysis techniques. Learn
details on KVL and KCL with examples.
Basic Circuit Design: Start designing your circuits! Explore different types
of circuits such as voltage dividers, amplifiers, and filters. Get hands-on
experience designing and building your own basic circuits.
Semiconductor Devices: Uncover the magic of semiconductors. Learn
about diodes and transistors, their types, functions, and applications in
modern electronics. Learn details on diode, diode circuits, transistor and
transistor circuits, biasing methods, etc.
Digital Electronics: Enter the world of digital logic gates and binary
arithmetic. Understand how digital circuits form the backbone of computers
and modern technology.
Practical Applications: Discover how Electronics concepts apply to real-
world applications. From simple gadgets to complex systems, you'll see
how electronics is everywhere.
Troubleshooting and Repair: Equip yourself with essential
troubleshooting skills to identify and fix common electronic problems.
Learn how to use multimeters and oscilloscopes effectively.
Safety in Electronics: Safety is paramount when working with electronics.
We'll cover best practices for safely handling electronic components and
equipment.
This book contains valuable information such as reference materials,
component datasheets, circuit diagrams, and additional resources such as
links to video lectures on our YouTube channel, VPACHKAWADE
Research Center
Link/s to all related YouTube lectures for easy references
https://www.youtube.com/playlist?list=PLWQzGc-sfVeWR-
UJ5xlcJMLVevHI3SA2V
Chapter 1 Introduction
Electronics and its importance in today’s world
Figure 1‑1 Emerging application areas of electronics and the research
being carried out at VPACHKAWADE Research center. For details visit,
www.vpachkawade.com
Electronics plays a pivotal role in today's world, shaping nearly every
aspect of our lives. From the smartphones we carry in our pockets to the
advanced machinery powering industries, electronics are the heartbeat of
modern society. They have revolutionized communication, enabling
instant global connections. Additionally, electronics drive innovations in
healthcare, transportation, entertainment, and countless other domains,
improving efficiency and quality of life. Moreover, they are at the core
of renewable energy solutions, contributing to a more sustainable future.
In essence, electronics are the foundation upon which our modern world
is built, and their importance continues to grow as technology evolves.
Figure 1‑1 shows several possible emerging areas of electronics
technology and applications.
Modern surface mount electronics components | components |
SMD
Figure 1‑2 showcases an electronics development board, specifically a
printed circuit board (PCB). The PCB hosts an array of modern
electronic components, each thoughtfully placed on this circuit board.
Within this assembly, you can see a series of surface-mount resistors.
Additionally, there's a surface-mount capacitor. There are also two
distinct components, possibly LEDs or other similar devices.
Furthermore, there's an integrated chip, commonly referred to as an IC.
You can identify this IC by referring to the data number on its package,
and you can look up its datasheet on Google to learn more about its
specific functionality.
Figure 1‑2 Modern surface-mount electronic components on a printed
circuit board (PCB), with a large IC at the top [60].
Similarly, the codes printed on the bodies of these resistors, as well as
other components, provide crucial information about their values. You
can consult the manufacturer's datasheets to gain a deeper understanding
of these components.
One noticeable feature is a regulator IC present on the board, with an
identifiable code on its body. A simple Google search can help you
determine the type of component or IC that this represents.
This is an example of what modern electronic development boards look
like, featuring miniaturized versions of electronic components and
integrated circuits.
Basic electronic components and their symbols
In the world of electronics, there are several fundamental components
that serve as the building blocks for all electronic devices.
Understanding these components and their symbols is essential for
anyone diving into the fascinating realm of electronics. Figure 1‑3 shows
various electronic components used in modern electronics. Electronic
components are fundamental parts used to build electronic circuits and
devices. They can be categorized into two main types: passive and active
components. Passive Components are basic electronic elements that do
not require an external power source to function. They include: resistors,
capacitors, inductors, and diodes. Active Components need an external
power source to operate and control electrical current. They include:
diodes, transistors, integrated circuits such as Op-Amps, voltage
regulators, LED (Light Emitting Diode) etc.
Figure 1‑3 Various electronic components [1]
Figure 1‑4 shows range of electronic components symbols used in
common electronics applications. We shall also explore details of each
of these components in the following sections.
Resistor (R):
A resistor limits the flow of electric current in a circuit. The symbol for a
resistor is the Greek letter omega (Ω), and its value is measured in ohms
(Ω). Resistors come in various sizes and shapes, and their job is to resist
or reduce the flow of electricity.
Resistor Capacitor Inductor Diode
Transistor ICs
LED [2]
Figure 1‑4 Various electronic components widely used in electronic -
symbols
Capacitor (C):
Imagine a capacitor as a temporary battery. It stores electrical energy and
releases it when needed. Capacitors are used for various purposes, from
smoothing out power supplies to timing circuits.
Inductor (L):
An inductor is like a coil of wire. It stores electrical energy in a magnetic
field. Inductors resist changes in current flow, making them valuable in
filters and energy storage.
Diode (D):
A diode allows current to flow in only one direction. They play a critical
role in rectifying AC (alternating current) to DC (direct current) and
protecting circuits from reverse voltage.
Transistor (NPN, PNP):
Transistors are like electronic switches that can amplify or control
current. There are different types, such as NPN and PNP. They are
essential for amplifying signals and building logic gates in digital
circuits.
Integrated Circuit (IC):
An integrated circuit is like a tiny city of electronic components packed
into one chip. ICs contain multiple electronic components, like
transistors and resistors, all working together to perform various
functions. They're the brains behind modern electronic devices.
These are the basic electronic components you'll encounter frequently in
electronic circuits. Understanding their symbols and functions is
important in electronics. Each component has a unique role in shaping
the behavior of circuits, making them the foundation of all electronic
devices.
Light Emitting diode (LED):
An LED, or Light Emitting Diode, is a tiny but powerful electronic
device that shines with vibrant colored light. LEDs work by allowing an
electric current to flow through them, which triggers them to glow. They
are commonly used in various everyday things like digital displays,
indicator lights on appliances, and even in dazzling light shows. LEDs
are favored for their energy efficiency, lasting a long time, and their
ability to produce bright and colorful light. Whether it's the indicator on
your phone or the colorful decorations during the holiday season, LEDs
make our world a little brighter and more colorful.
Overview of electronic circuits
Figure 1‑5 Range of electronic components used to build circuits, and
other useful functions of electronics hardware
In our exploration of electronics, it's essential to start with an overview
of electronic circuits. Imagine circuits as the roadways of the electronic
world, guiding electricity to perform specific tasks. These circuits are
built from basic components like resistors, capacitors, diodes, and
transistors, each with its unique job. We use symbols to represent these
components on paper. Power sources, such as batteries, provide the
energy needed to make circuits work. There are different types of
circuits, some for carrying continuous signals (analog) and others for
processing digital information (digital). Circuit diagrams help us
understand how everything is connected. In our journey, we'll explore
various circuit configurations, uncover their functions, and see how they
power the devices that make our lives easier and more exciting.
Chapter 2 Basic Electronic Components
Resistors
We have learned earlier elementary knowledge about what resistor is and
how it works, its relationship with voltage and current through ohm’s law.
Resistors are widely used in design of electronic circuits and systems,
where, they are used for voltage division, current limiting, and signal
conditioning.
Types of Resistors
a) Fixed Resistors: There are two types here.
Carbon Composition Resistors:
Carbon composition resistors are traditional passive electronic components
known for their simplicity and durability (see Figure 2‑1). These resistors
are constructed using a mixture of finely ground carbon particles and a
ceramic or clay binder. This combination forms a resistive element that
hinders the flow of electric current. Carbon composition resistors have been
widely used in various electronic applications for decades. They are
characterized by their ruggedness, high voltage-handling capabilities, and
resistance to mechanical stress and vibration. However, they may have
limitations in terms of precision and temperature stability compared to
more modern resistor types. Despite their declining use in contemporary
electronics due to advancements in resistor technology, carbon composition
resistors still find application in specific areas where their unique
characteristics are advantageous, making them an essential component in
vintage and niche electronic projects.
(a) (b)
Figure 2‑1 Various types of through-hole resistors used in electronics along
with their characteristics, (a) Carbon composition resistors [3], and (b)
Metal Film Resistor
Metal Film Resistors:
Metal film resistors are a common type of fixed resistor used in electronic
circuits, see Figure 2‑1. They are known for their precision, stability, and
low noise characteristics. These resistors are constructed by depositing a
thin layer of metal, often nickel-chromium or similar alloys, onto a ceramic
substrate. This thin metal layer serves as the resistive element. Metal film
resistors come in various resistance values, from fractions of an ohm to
several megaohms, and they typically have tight tolerances, such as ±0.1%
or ±1%, ensuring that their resistance values are very close to the specified
values. These resistors are commonly used in applications where accurate
and stable resistance values are crucial, such as in precision amplifiers,
analog signal processing, audio equipment, and measurement instruments.
Variable Resistors (Potentiometers):
Potentiometer [4]
[5]
Figure 2‑2 Different types of potentiometer
Their low noise and low temperature coefficient make them ideal for
applications requiring minimal interference and high precision. Metal film
resistors have become a fundamental component in modern electronics,
contributing to the reliability and performance of a wide range of electronic
devices and circuits. Variable resistors, commonly known as potentiometers
or "pots," are versatile electronic components that allow users to adjust
resistance within a specific range. The primary function of a potentiometer
is to offer variable resistance. Potentiometers consist of a resistive element,
a wiper (contact), and a mechanical shaft or knob (see Figure 2‑2). The
resistive element is a long, coiled or flat strip of resistive material (typically
carbon or conductive plastic) that is mounted inside the potentiometer
housing. By turning a knob or shaft, users can change the resistance within
a predefined range. The wiper is a movable contact that slides along the
resistive element as the shaft or knob is turned. The wiper's position
determines the effective resistance between the wiper terminal and the
other two terminals. Terminals are provided for connecting the
potentiometer in a circuit. Most potentiometers have three terminals: one
for the input voltage (fixed end), one for the variable output (wiper), and
one for the reference (the other end of the resistive element).
Potentiometers are available in various resistance values and can cover a
wide range, from tens of ohms to several megaohms. Typical potentiometer
tolerances range from ±5% to ±20%, depending on the precision required
for the application.
Resistor color code
Figure 2‑3 Resistor color code for knowing the values of resistors
The resistor color code (Figure 2‑3) is used to visually represent the
resistance value, tolerance, and sometimes the temperature coefficient of a
resistor. It consists of color bands or rings that are typically found on
resistors, but it's not commonly used for surface mount resistors (SMD),
which often have numerical codes.
Most resistors have either four or five colored bands. Each color
corresponds to a digit or a multiplier
1. First Band (Significant Digit): The color of the first band
represents the first digit of the resistance value.
2. Second Band (Second Significant Digit): The color of the
second band represents the second digit of the resistance value.
3. Third Band (Multiplier): The color of the third band indicates
the multiplier by which to multiply the significant digits to get
the resistance value. For example, if the third band is red (2), you
multiply the significant digits by 102 (100).
4. Fourth Band (Tolerance): If there is a fourth band, it represents
the tolerance of the resistor. The color codes for tolerance are:
Gold (±5%)
Silver (±10%)
No Color (±20%)
Brown (±1%) (less common)
5. Fifth Band (Temperature Coefficient, Optional): Some
precision resistors may have a fifth band that represents the
temperature coefficient. Common color codes for temperature
coefficient are:
Brown (100 ppm/°C)
Red (50 ppm/°C)
Orange (15 ppm/°C)
Yellow (25 ppm/°C)
Blue (10 ppm/°C)
Reading the Value: To read the resistance value, look at the first two bands
for the significant digits, the third band for the multiplier, and the fourth
band (if present) for the tolerance.
For example, if you see a resistor with bands: Yellow, Violet, Red,
Gold, it can be decoded as follows:
Significant digits: Yellow (4) and Violet (7)
Multiplier: Red (2)
Tolerance: Gold (±5%)
So, the resistance value is 47 x 102 ohms, or 4.7 kΩ with a
tolerance of ±5%. Now, verify the values of the resistors given in
Figure 2‑3.
What is Surface Mount Device
(SMD) resistor?
Resistors (SMD) on Printed circuit 100 ohm SMD 100 ohm
board [6] SMD [7]
Figure 2‑4 SMD resistors soldered on board and parts available in the
market
Surface Mount Device (SMD) resistors are compact and versatile electronic
components designed for modern electronics manufacturing and
miniaturized circuits (Figure 2‑4). These resistors are characterized by their
small, surface-mountable package that allows them to be soldered directly
onto printed circuit boards (PCBs). SMD resistors come in various shapes
and sizes, making them suitable for a wide range of applications, from
consumer electronics to industrial equipment. Their advantages include
space-saving design, excellent thermal performance, and compatibility with
automated assembly processes. SMD resistors play a crucial role in
contemporary electronics, contributing to the miniaturization and efficiency
of electronic devices.
SMD resistors typically have numerical or alphanumeric codes printed on
their surface (Figure 2‑4).
How to read the SMD Code: You should find a numerical or
alphanumeric code printed on SMD resistor surface. This code can vary in
format and length depending on the manufacturer and series. Different
manufacturers may use slightly different code formats, but generally, the
code represents the resistance value in ohms and possibly other information
like tolerance and package size.
To determine the specific resistance value of the SMD resistor, you'll need
to refer to the datasheet or documentation provided by the manufacturer.
The datasheet will include a key or table that explains how to decode the
SMD code for that particular series of resistors.
There are also online calculators and databases available that can help you
decode SMD resistor codes [8] [9]. You can enter the code, and these tools
will provide you with the resistance value and other relevant information.
See the links given below:
https://www.utmel.com/tools/smd-resistor-code-calculator?id=33
https://www.digikey.in/en/resources/conversion-calculators/conversion-
calculator-smd-resistor-code
Capacitors
In section Error! Reference source not found., we learned what capacitor
is. When a voltage is applied across the two plates of a capacitor, it stores
electrical charge on its plates. When the voltage is removed or reduced, the
capacitor discharges and releases the stored energy.
What is a Capacitor?
The basic structure of a capacitor consists of two conductive plates
separated by a dielectric material (Figure 2‑5). The conductive plates are
typically made of metal, such as aluminum, tantalum, or copper. The
dielectric material is a non-conductive substance placed between the two
conductive plates. It is usually made of materials like ceramics, plastics,
paper, or electrolytic materials. The dielectric material serves as an
insulator, preventing direct electrical contact between the plates. This
insulation is essential because it allows the capacitor to store electrical
charge without short-circuiting.
Parallel plate capacitor Capacitor symbols
[10]
Figure 2‑5 Structure of a parallel plate capacitor, showing tow plates
separated by air as a dielectric
Capacitance is defined as the measure of a capacitor's ability to store
electrical charge. It quantifies the amount of electric charge a capacitor can
hold for a given voltage across its terminals. The formula for capacitance
(C) of a capacitor is:
C=Q/V
Where:
C is the capacitance in farads (F).
Q is the electric charge stored on the capacitor's plates in
coulombs (C).
V is the voltage across the capacitor's plates in volts (V).
This formula illustrates that the capacitance of a capacitor is directly
proportional to the amount of electric charge it can store for a given
voltage. In other words, the higher the capacitance, the more charge the
capacitor can hold for a given voltage, making it a more effective energy
storage device. The unit of capacitance is the farad (F). One farad is
defined as the capacitance of a capacitor that stores one coulomb (C) of
electric charge when a voltage of one volt (V) is applied across its
terminals:
1 farad (F) = 1 coulomb (C) / 1 volt (V).
Capacitance values are specified on capacitors using their appropriate unit
(µF, nF, pF, or F). Figure 2‑5 also shows symbols used for fixed, polarized
and variable capacitors, discussed below.
Types of capacitors
Figure 2‑6 shows various types of capacitors used in electronics applications.
Each are described below.
1. Electrolytic Capacitors:
Electrolytic capacitors are polarized, meaning they have a positive and
negative terminal. They typically have high capacitance values and are
available in
1000μF
Electrolytic
Capacitor | Free
SVG
Types of capacitors [11] Electrolytic Ceramic capacitor [12]
capacitor
SMD Ceramic capacitor tantalum capacitor Film capacitor [12]
[13] [14]
Figure 2‑6 Types of capacitor used in electronics
cylindrical or rectangular shapes. Aluminum and tantalum are common
materials for the electrodes. Electrolytic capacitors are used for high-
capacitance, low-frequency applications, such as power supply filtering and
audio signal coupling. They are also used in energy storage for applications
like camera flashes.
2. Ceramic Capacitors:
Ceramic capacitors are small and come in a wide range of capacitance
values. They are made from ceramic materials with metallic electrodes.
They are non-polarized. They are suitable for high-frequency applications,
decoupling, bypassing, and noise suppression in digital and RF circuits.
3. Tantalum Capacitors:
Tantalum capacitors are polarized and offer a high level of capacitance in a
small package. They are known for their stability and reliability. Tantalum
capacitors are used in power supply filtering, coupling, and low-frequency
timing circuits. They are common in portable electronics and
telecommunications equipment.
4. Film Capacitors:
Film capacitors are non-polarized and come in various dielectric materials,
including polyester, polypropylene, and polyethylene. They are available in
a range of capacitance values. Film capacitors are used in audio
applications for signal coupling and filtering. They are also suitable for
high-voltage and high-frequency applications.
5. Variable Capacitors:
Variable capacitors have a variable capacitance that can be adjusted
manually or electronically. They consist of plates that overlap, and the
capacitance changes as the plates move. Variable capacitors are used in
tuning circuits for radios, televisions, and antennas. They allow precise
tuning of resonant circuits.
6. Mica Capacitors:
Mica capacitors use mica as the dielectric material and have excellent
stability and low loss. Mica capacitors are used in high-frequency
applications, such as radio transmitters and receivers, and in precision timing
circuits.
7. Glass Capacitors:
Glass capacitors use glass as the dielectric material and are known for their
stability and low temperature coefficient. Glass capacitors are used in
precision timing, tuning, and oscillator circuits.
8. Variable Capacitors:
The capacitance value of variable capacitors is often indicated on
the knob or adjustment mechanism rather than on the capacitor
body. It may be represented in picofarads (pF) or another
appropriate unit.
Reading a capacitor markings and
codes
Refer to Figure 2‑6.
1. Electrolytic Capacitors:
Electrolytic capacitors often have a numeric value followed by a
unit, such as µF (microfarads) or mF (millifarads). For example,
"47µF" means 47 microfarads. Larger capacitors may use codes
to represent capacitance values, where a numeric code
corresponds to a specific capacitance value.
2. Ceramic Capacitors:
Ceramic capacitors often use a three-digit code to represent their
capacitance value. The first two digits indicate significant
figures, and the third digit indicates the number of zeros to add.
For example, "104" represents 100,000 pF or 0.1 µF. Figure 2‑6
shows a 220pF ceramic capacitor.
3. Tantalum Capacitors:
Tantalum capacitors typically have a numeric value followed by
a unit, such as µF or mF. The value is straightforward to read, as
it directly represents the capacitance in microfarads or
millifarads.
4. Film Capacitors:
Film capacitors usually have their capacitance value explicitly
marked as a numeric value followed by µF, nF (nanofarads), or
pF (picofarads).
The voltage rating of a capacitor is usually indicated as a voltage value (in
volts) preceded by a "V" or "WV" (working voltage). For example, "25V"
means a voltage rating of 25 volts.
How does capacitor charge and
discharge?
File:RC circuit MK ch dch.png - Wikimedia Commons
Figure 2‑7 Charging and discharging process of a capacitor in a simple RC
circuit [12]
Capacitors charge and discharge in response to changes in voltage
according to certain principles. Let's explore how capacitors behave during
charging and discharging: (Figure 2‑7)
1. Charging of Capacitors:
When a voltage source (e.g., a battery) is connected across the
terminals of a capacitor, the capacitor begins to charge.
Electrons from the negative terminal of the voltage source flow
onto the capacitor's plate connected to the negative terminal,
while electrons from the other plate flow away to the positive
terminal of the source.
As the process continues, the voltage across the capacitor
increases, and the rate of charge accumulation decreases over
time.
In an ideal scenario with no resistance in the circuit, a capacitor
will charge to the source voltage in a theoretically infinite
amount of time.
2. Discharging of Capacitors:
When a charged capacitor is disconnected from its voltage source
and connected to a circuit with a lower voltage or a resistor, it
begins to discharge.
Electrons flow from one plate of the capacitor through the circuit
(or resistor) to the other plate, gradually reducing the voltage
across the capacitor.
Similar to charging, the rate of discharge decreases over time.
In an ideal scenario with no resistance in the circuit, a capacitor
will take an infinite amount of time to fully discharge.
The RC Time Constant (τ):
In practical circuits, there is often some resistance in the circuit due to the
wires, connections, and sometimes intentional resistors (Figure 2‑7). This
resistance affects the rate at which a capacitor charges and discharges. The
time it takes for a capacitor to charge to approximately 63.2% of its final
voltage (or discharge to 36.8% of its initial voltage) in the presence of
resistance is defined by the RC time constant (τ). The RC time constant is
calculated using the following formula:
τ=RxC
Where:
τ (tau) is the time constant in seconds (s).
R is the resistance in ohms (Ω) in the circuit.
C is the capacitance in farads (F) of the capacitor.
The RC time constant characterizes the time it takes for a capacitor to reach
about 63.2% of its final voltage during charging or discharge. It also
indicates how quickly a capacitor responds to changes in voltage. A smaller
τ results in a faster response, while a larger τ leads to a slower response.
Inductors
let's understand what an inductor is. An inductor is a two-terminal passive
electrical component that stores energy in the form of a magnetic field
(Figure 2‑8). When an electric current passes through the inductor, it
produces a surrounding magnetic field. If the current varies with respect to
time, it creates a time-varying magnetic field around it. This phenomenon
is known as electromagnetic induction, based on Faraday's law. As a result
of this action, an electromotive force (EMF), or voltage, is induced across
the two terminals of the inductor. The polarity of this EMF opposes the
direction of the current that produces it. Mathematically, the voltage across
the inductor is equal to the inductance (L) multiplied by the rate of change
of the current flowing through it.
File:Inductor Symbols.jpg -
Wikimedia Commons
Inductor [15] Symbols [16]
Figure 2‑8 Inductor and its symbols
In other words, the value of the inductance, denoted as L, can be defined as
the ratio of the voltage across it divided by the rate of change of current,
which is expressed as dI/dt. Typically, inductors are made by winding
insulated wire into a coil, and there are various types of inductors available,
with values ranging from a few mH to as high as 20 Henry, or even in the
mH range for smaller values. Mathematically, an inductance (L) is defined
as the ratio of magnetic flux (ΦB) to current (I), expressed as L = ΦB / I.
When an electromotive force (EMF) is generated across the inductor,
represented as VL, it is precisely the rate of change of magnetic flux with a
negative sign, VL = -dΦB/dt. Importantly, magnetic flux (ΦB) is
fundamentally a product of inductance (L) and current (I), ΦB = L * I. By
substituting this value into the expression for VL, we arrive at the final
mathematical representation: VL = -L * (dI/dt). These equations offer a
mathematical framework for understanding how inductors store energy in
magnetic fields and respond to changes in current, ultimately producing
induced EMFs in electrical circuits.
Chapter 3 Electronic Fundamentals
Alternating Current (AC) and Direct Current (DC) in
Electricity
Figure 3‑1 AC and DC, The horizontal axis measures time (it also
represents zero voltage/current); the vertical, current or voltage [17].
Alternating Current (AC) and Direct Current (DC) are two fundamental
types of electric currents with distinct properties and applications. AC is
characterized by its continuous change in direction, regularly shifting back
and forth. It is the form of electricity commonly used for delivering power
to homes and businesses. When you plug in your kitchen appliances,
televisions, fans, or electric lamps, they typically run on AC. The AC
waveform is often represented as a sine wave, which is a smooth, wavy
line. In contrast, DC flows steadily in one direction, as seen in batteries or
the power source for many electronic devices. While AC is versatile and
efficient for long-distance power transmission, DC is more stable and
reliable for electronic circuits. Both types have their unique strengths and
are used in various applications, making them essential components of our
modern electrical systems.
Voltage, current, and resistance
Voltage, current, and resistance are the fundamental building blocks of
electronics, and understanding these concepts is key to understand
electronic circuits.
Voltage (V):
Voltage, often denoted as "V," is a fundamental electrical quantity in
electronics. It represents the electric potential difference between two
points in an electrical circuit. In simpler terms, voltage can be thought of as
the "electrical pressure" that pushes electric charge (usually in the form of
electrons) through a conductor like a wire or an electronic component.
Figure 3‑2 shows symbol of a voltage source, various sources and
checkpoints where voltage is available. It also shows an application circuit
where voltage can be applied and measured. It also shows electronic power
supply that provides voltage to other circuits.
High voltage electric station High voltage Voltage divider in
symbol electrical circuit
Battery Voltage source unit Power supply unit
Symbol of a voltage
Figure 3‑2 Sources of voltage, symbol, circuit application and electronic
unit
Facts about voltage:
In an electrical circuit, voltage provides the force that moves
electric charge (current) from one point to another.
Voltage can be supplied by various sources, including
batteries, power supplies, generators, and electrical outlets.
These sources create a potential difference, causing electric
charge to flow when a circuit is completed.
Voltage is always measured relative to a reference point.
When we say an object has a voltage of, for example, 5 volts,
it means it has 5 volts more (or less) electrical potential
energy than the chosen reference point.
Voltage levels can signal "on" or "off" states in digital
circuits and control the brightness of a light-emitting diode
(LED).
Kirchhoff's Voltage Law (KVL) (discussed later) is a
fundamental principle in circuit analysis. It states that the
sum of the voltages around any closed loop in a circuit is
equal to zero, emphasizing the conservation of energy in
electrical circuits.
Voltage can pose electrical hazards, so it's essential to handle
electrical equipment and circuits with care. High-voltage
circuits can cause electrical shocks or damage to
components.
Current (I):
Current, denoted as "I," is a fundamental electrical quantity in
electronics and represents the flow of electric charge through a
conductor, such as a wire or a circuit. It is the rate at which electric
charge moves past a given point in a circuit and is measured in amperes
(A). See Figure 3‑3 for facts explained below.
Facts about current:
Current is the flow of electric charge, typically in the form of
electrons, through a conductive pathway.
Current is measured in amperes (A). One ampere is defined
as one coulomb of charge passing through a conductor per
second.
The direction of current flow is conventionally defined as the
direction in which positive charges would move. In reality,
electrons, which are negatively charged, flow in the opposite
direction.
Direct Current (DC): In DC circuits, current flows steadily in
one direction.
Alternating Current (AC): In AC circuits, current reverses
direction periodically, typically in a sinusoidal waveform.
AC is commonly used in household electricity.
Ohm's Law (explained below) is a fundamental principle in
electronics, it relates voltage (V), current (I), and resistance
(R) in a circuit. It is expressed as V = I x R, where voltage is
directly proportional to current and resistance.
The principle of current conservation states that in a closed
circuit, the total current entering a point is equal to the total
current leaving that point. This principle ensures that charge
is conserved within the circuit.
High current levels can pose electrical hazards, such as
electrical shocks and circuit overheating. Proper safety
precautions are necessary when working with electrical
currents.
Current can be provided by various sources, including
batteries, generators, and power supplies. These sources
create a potential difference (voltage) that drives current
through a circuit.
Charge flow via conductor AC symbol
Conventional current flow [18] Practical Voltage and a
current sources [19]
Figure 3‑3 symbol for a current, flow of charges, symbol for a AC
source, direction of current and sources of current (that produces a
voltage ) and voltage (that produces a current when load is connected)
Resistance (R):
Resistance, often denoted as "R," is a fundamental electrical property
that describes the opposition to the flow of electric current in a
conductor, such as
Resistance of a material [20] Ohm’s law depicting relation
between voltage, current and
resistance [21]
Resistor colour code
Figure 3‑4 Resistance of a material, ohm's law showing relation between
current, voltage and resistance and color code of a resistor
a wire, component, or circuit. It is a crucial concept in electronics and is
measured in ohms (Ω).
Facts about resistance:
Resistance represents the degree to which a material or
component resists the flow of electric current.
Resistance is measured in ohms (Ω). One ohm is defined as
the resistance that allows a current of one ampere (A) to flow
when a voltage of one volt (V) is applied across it.
Different materials have different inherent resistances. For
example, metals like copper have low resistance and are good
conductors, while materials like rubber have high resistance
and are insulators.
Longer conductors generally have higher resistance. A wider
conductor has lower resistance.
The resistance of some materials, especially semiconductors,
can change with temperature.
Ohm's Law relates voltage (V), current (I), and resistance (R) in a
circuit. It is expressed as V = I x R, indicating that voltage is
directly proportional to current and resistance.
When current flows through a resistor, it generates heat due
to the resistance. The power dissipated as heat in a resistor is
calculated using the formula P = V2 / R, where P is power, V
is voltage, and R is resistance.
To identify the resistance value of a resistor, a color code system is often
used, where different colored bands on the resistor correspond to specific
resistance values.
Resistance is a fundamental property in various electronic
applications, such as in voltage dividers, filters, and current-
limiting circuits.
Some components, like potentiometers and variable resistors, allow the
resistance value to be adjusted, making them useful for applications
where resistance needs to be variable. See Figure 3‑4.
Ohm's Law and how to apply it
Ohm's Law is a fundamental principle in electronics, named after the
German physicist Georg Simon Ohm. It describes the relationship
between voltage (V), current (I), and resistance (R) in an electrical
circuit. See Figure 3‑4. Ohm's Law is expressed mathematically as V = I
x R, where:
V represents voltage in volts (V).
I represents current in amperes (A).
R represents resistance in ohms (Ω).
Ohm's Law tells us that the voltage across a component in a circuit
is directly proportional to the current flowing through it. If you increase
the voltage, the current increases proportionally, assuming the resistance
remains constant.
Ohm's Law allows you to calculate any one of the three variables
(voltage, current, or resistance) if you know the values of the other two.
This is useful in designing and analyzing circuits. For example, if you
have a resistor with a known resistance (R) and a known current (I)
passing through it, you can find the voltage drop (V) across the resistor
using V = I x R.
Figure 3‑5 Voltage divider circuit
Ohm's Law is fundamental in designing voltage divider circuits (see
voltage divider circuit in Figure 3‑5. By connecting resistors in series,
you can create a voltage divider that provides a specific voltage output
relative to the input voltage. This is commonly used in applications like
adjusting brightness in LED displays or setting reference voltages.
Ohm's Law is essential for current limiting in circuits to protect
components or prevent excessive current flow. By choosing the
appropriate resistance value, you can limit the current passing through a
component to a safe level. Engineers and hobbyists use Ohm's Law
extensively when analyzing circuits. It helps determine how voltages and
currents behave in different parts of a circuit, making it a valuable tool
for troubleshooting and optimizing circuit performance.
Ohm's Law also plays a role in power calculations. You can use it
to find the power dissipated in a resistor using the formula P = V x I or P
= I2 x R. These equations are crucial for selecting components that can
handle the power generated in a circuit.
Ohm's Law is foundational in a wide range of electronic
applications, including power supplies, amplifiers, lighting circuits, and
more. It ensures that circuits operate as intended and within safe limits.
Electrical power and energy
Understanding electrical power and energy is essential in electronics and
electrical engineering. These concepts are fundamental for designing
circuits, calculating energy consumption, and optimizing the efficiency
of electrical systems. (see High voltage electric station in Error!
Reference source not found.)
Electrical Power (P):
Electrical power, denoted as "P," is the rate at which electrical energy is
consumed, generated, or transferred in an electrical circuit. It quantifies
how quickly electrical work is done.
The formula for electrical power is given by P = V x I, where:
P represents power in watts (W).
V represents voltage in volts (V).
I represents current in amperes (A).
Facts:
Power is measured in watts (W), and it indicates the rate of
energy transfer or consumption per unit of time.
In a circuit, power represents the ability to do work, such as
lighting a bulb or running a motor.
High-power devices consume more electrical energy per unit
of time than low-power devices.
Power is a crucial factor in designing circuits, selecting
components, and calculating power losses.
Power ratings are often used to specify the capacity or
performance of electrical devices. For example, a 100W light
bulb consumes electrical power at a rate of 100 watts when
lit.
Engineers use power to assess the efficiency of electrical
systems. By comparing input and output power, they can
determine how effectively energy is converted and used.
Electrical Energy (E):
Electrical energy, denoted as "E," is the total amount of work done or
energy consumed by an electrical device or system over a period of time.
It is measured in watt-hours (Wh) or kilowatt-hours (kWh).
Electrical energy can be calculated using the formula E = P x t, where:
E represents energy in watt-hours (Wh) or kilowatt-hours (kWh).
P represents power in watts (W).
t represents time in hours (h).
Facts:
Electrical energy is a cumulative measure of power
consumption over time. It is what you are billed for on your
electricity bill.
1 kilowatt-hour (kWh) is equivalent to 1,000 watt-hours
(Wh). It's a common unit for measuring electrical energy in
residential and commercial settings.
Electrical energy is used to assess the long-term cost of
operating electrical devices and systems.
Assessing energy consumption helps evaluate the
environmental impact of electrical systems, as it relates to
carbon emissions and sustainability efforts.
Engineers use energy calculations to design efficient systems
and optimize energy usage in various applications, from
renewable energy systems to electric vehicles.
Kirchhoff's laws as circuit analysis techniques
In this module, we will explore a fundamental law in network analysis
known as Kirchhoff's law, named after the German scientist Mr.
Kirchhoff. There are two Kirchhoff's laws: Kirchhoff's current law
(KCL) and Kirchhoff's voltage law (KVL). First, we will focus on
Kirchhoff's current law, often abbreviated as KCL.
Kirchhoff's current law (KCL)
Consider a network or circuit, as shown (Figure 3‑6), with currents
labeled as I1, I2, I3, and I4, each with its specified direction. The central
point, represented as a black dot, is known as a node or junction within
the network.
Kirchhoff's current law (KCL) states that the algebraic sum of the
currents entering a specific node is equal to the sum of the currents
exiting the same node. In other words, the sum of currents flowing into a
node equals the sum of currents flowing out of that node.
Figure 3‑6 Kirchhoff's current law [22]
For instance, at this node, we observe that I2 and I3 are entering.
Therefore, we can express KCL as:
I2 + I3 = I1 + I4.
This equation signifies that the sum of currents entering (I2 and I3) equals
the sum of currents exiting (I1 and I4) the node.
Alternatively, we can express KCL as the algebraic sum of all currents in
a network being equal to zero:
I2 + I3 - I1 - I4 = 0.
This equation reinforces that in a network or circuit with conductors, the
algebraic sum of all currents is always zero, representing the
conservation of electric charge.
Kirchhoff's current law (KCL) is applicable to both linear and non-linear
circuits, making it a fundamental principle in network analysis.
Kirchhoff's voltage law (KVL)
Figure 3‑7 Kirchhoff’s voltage law [22]
In this module, we are going to understand the second law from
Kirchhoff, called Kirchhoff's voltage law, or KVL for short. Imagine you
have a network or a circuit as shown in the figure. It consists of a voltage
source known as V4, resistors R1, R2, R3, and another resistor R5.
Additionally, you have black dots labeled as a, b, c, and d, which are
referred to as nodes or junctions within the circuit.
In the network shown in Figure 3‑7, we have voltage drops across R1
(indicated as V1), R2 (indicated as V2), R3 (indicated as V3), and R5
(indicated as V5) due to the current flowing through these resistors as a
result of the voltage source. According to Ohm's law, these voltage drops
occur.
Since R3 and R5 are connected in parallel, as both of their ends are
connected to the same nodes or junctions, the voltage drop across each
of them is the same. In other words, V3 and V5 are equal.
Our focus in this module is to explain Kirchhoff's voltage law, or KVL.
This law states that the algebraic sum of all the voltages around a closed
loop is equal to 0. Mathematically, we can express KVL as:
V1 + V2 + V3 + V4 = 0.
This equation defines KVL, where V3 and V5 are considered equal.
KVL can be applied to any circuit, whether it is linear or non-linear,
consists of only passive elements, active elements, or a combination of
both.
Example 1
Figure 3‑8 Example 1 circuit [23]
In this network (Figure 3‑8), we can indeed apply Kirchhoff's Voltage
Law (KVL). KVL states that the algebraic sum of all the voltages around
a closed loop is equal to zero. Allow me to explain how this applies to
the given linear circuit, which consists of a voltage source and two
resistors in series.
Firstly, we have the voltage source, denoted as VS. Due to this voltage
source, a current, denoted as I, flows through the circuit. This current
passes through both R1 and R2 and returns to the negative terminal of the
source. The voltage drop across R2 due to the current I is denoted as
VR2, and the voltage drop across R1 is denoted as VR1.
Now, let's consider the voltages present in the circuit: VS, VR2, and VR1.
According to KVL, we can express it as:
VS + VR2 + VR1 = 0.
This equation represents Kirchhoff's Voltage Law, stating that the sum of
all voltages around a closed loop is equal to zero.
Example 2
Figure 3‑9 Example circuit 2 [24]
In this circuit (Figure 3‑9), we are going to apply Kirchhoff's Current
Law (KCL). You can see that there are currents denoted as i1, i2, i3, i4,
and i5. This specific point is referred to as a node or junction. It's
important to note that this point is called a node or junction.
KCL states that in a network, as shown in this figure, the sum of the
currents entering a node must equal the sum of the currents exiting the
same node.
For this node, we have i1 and i2 entering, so we can express this as:
i1 + i2
On the exiting side, we have i3, i4, and i5, therefore,
i1 + i2 = i3 + i4 + i5.
This equation demonstrates Kirchhoff's Current Law in action.
Series and parallel circuits
In electronics, there are two ways to connect components: series and
parallel. When components are in series (Figure 3‑10), they form a
single path for electric current, and the current is the same through each
component. In parallel, components have multiple paths, and the voltage
is the same across each one.
Figure 3‑10 A series circuit with a voltage source (such as a battery, or a
cell) and three resistance units [25]
Imagine a simple circuit with four light bulbs and a 12-volt battery. If the
bulbs are connected one after the other in a loop, they are in series. In
this case, each bulb only gets 3 volts, which might not make them glow.
However, if the bulbs are connected separately to the battery, they are in
parallel. Now, each bulb gets the full 12 volts, and they all glow.
The key difference is that in a series circuit, if one component fails, the
whole circuit breaks. But in a parallel circuit, if one component fails, the
others can still work. This is why parallel circuits are commonly used in
homes and devices to ensure that if one part stops working, the rest can
still function.
Series circuits
In a series circuit, the current is the same for all of the components, and
it can be calculated using the following equation:
I_total = I_1 = I_2 = ... = I_n
Where:
I_total is the total current in the series circuit.
I_1, I_2, ..., I_n are the currents through each individual component in the
series circuit.
In a series circuit, the total voltage (V_total) across the circuit is equal to
the sum of the voltages across each individual component. This can be
expressed using the equation:
V_total = V_1 + V_2 + ... + V_n
Where:
V_total is the total voltage in the series circuit.
V_1, V_2, ..., V_n are the voltages across each individual component in the
series circuit.
In a series circuit, the total resistance (R_total) is equal to the sum of the
resistances of the individual components. You can express this using the
equation:
Figure 3‑11 Resistance in series [25]
R_total = R_1 + R_2 + ... + R_n
Where:
R_total is the total resistance in the series circuit.
R_1, R_2, ..., R_n are the resistances of each individual component in the
series circuit.
Conductance (G) is the reciprocal of resistance in an electrical circuit.
In a series circuit, the total conductance (G_total) of pure resistances is
calculated by the sum of the reciprocals of the individual conductances.
The equation for calculating G_total in a series circuit is:
1/G_total = 1/G_1 + 1/G_2 + ... + 1/G_n
Where:
G_total is the total conductance in the series circuit.
G_1, G_2, ..., G_n are the conductances of each individual component in
the series circuit.
Conductance is typically measured in siemens (S), which is the
reciprocal of ohms (Ω). It represents how easily electric current can flow
through a component or circuit.
In a series circuit, inductors follow the same law as resistors and
capacitors.
Figure 3‑12 Inductance in series [25]
The total inductance (L_total) of non-coupled inductors in series is equal to
the sum of their individual inductances. The equation for calculating
L_total in a series circuit is:
L_total = L_1 + L_2 + ... + L_n
Where:
L_total is the total inductance in the series circuit.
L_1, L_2, ..., L_n are the inductances of each individual inductor in the
series circuit.
Inductance (L) is a property of inductors, and it is measured in henrys
(H). It represents the ability of an inductor to store electrical energy in
the form of a magnetic field when current flows through it. In a series
circuit, the total inductance is the sum of the individual inductances.
In a series circuit, capacitors also follow the same law as resistors and
inductors.
Figure 3‑13 Capacitance in series [25]
The total capacitance (C_total) of capacitors in series is equal to the
reciprocal of the sum of the reciprocals of their individual capacitances.
The equation for calculating C_total in a series circuit is:
1 / C_total = 1 / C_1 + 1 / C_2 + ... + 1 / C_n
Where:
C_total is the total capacitance in the series circuit.
C_1, C_2, ..., C_n are the capacitances of each individual
capacitor in the series circuit.
Capacitance (C) is a property of capacitors, and it is measured in farads
(F). It represents the ability of a capacitor to store electrical energy in the
form of an electric field when a voltage is applied across it. In a series
circuit, the total capacitance is calculated by adding the reciprocals of
the individual capacitances and taking the reciprocal of the sum.
When two or more switches are connected in series, they form a logical
AND gate in electronics. This means that for current to flow through the
circuit, all the switches must be closed (in the ON position). If any of the
switches in the series circuit is open (in the OFF position), it will
interrupt the flow of current through the entire circuit.
Cells and batteries are often connected in series to achieve higher
voltages. A battery is essentially a collection of electrochemical cells.
When these cells are connected in series, the total voltage of the battery
is equal to the sum of the individual cell voltages.
For example, let's consider a 12-volt car battery. It contains six 2-volt
cells connected in series. When these cells are connected this way, their
voltages add up to provide a total voltage of 12 volts for the battery. This
higher voltage is essential to power various electrical components in a
car.
In some cases, such as trucks, two 12-volt batteries may be connected in
series to create a 24-volt system. When batteries are connected in series,
their voltages are additive, which makes them suitable for applications
requiring higher voltage levels.
This series connection of cells or batteries is a common technique to
meet specific voltage requirements in various electrical and electronic
devices.
Example:
Figure 3‑14 Series connection of resistors [26]
Do you know how to determine the effective resistance of network in
Figure 3‑14? When I refer to effective resistance, I mean the total or
equivalent resistance from point A to point B. We're examining this
network configuration, which shows that resistors R1, R2, R3, R4, and
R5 are connected in series. Therefore, the total resistance from point A
to point B can be calculated as the sum of R1, R2, R3, R4, and R5. If
each of these resistors has a value of 1 kilo ohm, then the total resistance
would be 5 kilo ohms. This clearly demonstrates a series connection of
the resistors.
R_total = R1 + R2 + R3 + R4 + R5
In this case, if each resistor has a value of 1 kilo ohm, the total resistance
(R_total) would be:
R_total = 1 kΩ + 1 kΩ + 1 kΩ + 1 kΩ + 1 kΩ = 5 kΩ
Parallel circuits
In a parallel circuit, components are connected in multiple paths, and
each component has the same voltage across it, equal to the voltage
across the network. The current through the network is equal to the sum
of the currents through each component.
Voltage (V):
In a parallel circuit, the voltage (V) is the same across all components.
V_total = V1 = V2 = V3 = ... = Vn
Current (I):
The total current (Itotal) is the sum of the currents through each branch
or component in parallel.
Itotal = I1 + I2 + I3 + ... + In
(a) (b)
(c)
(d)
Figure 3‑15 parallel circuits for resistive, capacitor and inductor
network, (a to c) and final summary of series and parallel combination of
passive components in electrical network [27]
Figure 3‑15 shows parallel circuit for resistors, capacitors, inductors, and
finally also provides a summary of series and parallel circuits of the
passive components.
Resistance (R):
In a parallel circuit, the reciprocal of the equivalent resistance (1/R_total)
is the sum of the reciprocals of the individual resistances.
1 / R_total = 1 / R1 + 1 / R2 + 1 / R3 + ... + 1 / Rn
Capacitors (C):
In a parallel circuit, the total capacitance (C_total) is the sum of the
individual capacitances.
C_total = C1 + C2 + C3 + ... + Cn
Inductors (L):
For non-coupled inductors in parallel, the total inductance (L_total) is the
sum of the individual inductances.
Equation for Total Inductance (L_total):
L_total = L1 + L2 + L3 + ... + Ln
Switches:
In a parallel circuit, two or more switches in parallel form a logical OR.
The circuit carries current if at least one switch is closed.
Cells and Batteries:
In a parallel configuration, the voltage of the battery remains the same,
and the total current capacity is the sum of the individual cells.
Equation for Total Voltage (V_total):
V_total = V1 = V2 = V3 = ... = Vn
Example:
Determine the equivalent network of the circuit at its two open ends.
Figure 3‑16 Resistive network [27]
To determine the equivalent resistance of the network between its two
open points, let's break it down:
1. Resistors R1 and R2 are in parallel to each other because both
ends of R1 connect to both ends of R2. The equivalent
resistance (Rp) of two resistors in parallel can be calculated
using the formula:
1/Rp = 1/R1 + 1/R2
2. Now, resistor R3 is in series with the parallel combination of
R1 and R2. The equivalent resistance (Rs) of resistors in series
is simply the sum of their resistances:
Rs = R3 + Rp
So, the total equivalent resistance (R_total) of the network between the two
open points is:
R_total = R3 + (1 / (1/R1 + 1/R2))
If you have specific resistance values for R1, R2, and R3, you can
substitute those values into the equation to calculate the equivalent
resistance.
Voltage and current dividers
Voltage division
(a) (b) (c)
Figure 3‑17 voltage division
Let's understand how the voltage division circuit works. Imagine you
have two resistors, R1 and R2, connected in series (Figure 3‑17 a). You
apply an input voltage, Vin, to one end of R1. At the junction of R1 and
R2, you tap into the circuit and measure the output voltage, Vout.
Mathematically, Vout can be expressed as:
Vout = Vin × (R2 / (R1 + R2))
Essentially, we're measuring the voltage across resistor R2, so Vout is
also the voltage across R2, denoted as VR2.
Now, when you apply a voltage to this circuit, a current, I, flows from
the supply towards the ground. This current, I, flows through both R1
and R2 because they are connected in series. You can express the current
as Ir1 = Ir2 = I ( Figure 3‑17 b). So, the same current flows through both
R1 and R2 due to their series connection.
We can apply Ohm's law to calculate the voltage across R2. Using the
equation Vout = Vin × (R2 / (R1 + R2)), we can also express VR2 as
VR2 = I × R2.
Similarly, if you want to calculate the voltage across R1, you can use
VR1 = I × R1, which is a straightforward application of Ohm's law. If
you want to calculate the voltage across R1 using the voltage division
rule, it can be expressed as VR1 = Vin × (R1 / (R1 + R2)) (Figure 3‑17
c).
Comparing these equations, you'll notice that whether you calculate
Vout, VR2, or VR1, you select the resistor across which you want to
determine the voltage and place its resistance in the numerator. In the
denominator, you use the sum of all the resistors in the circuit.
This concept is known as the voltage division rule, and circuits like this
are called voltage dividers. We've calculated voltages and currents using
Ohm's law and the voltage division rule.
Example:
(a)
(b) (c) (d)
Figure 3‑18 Another example of voltage division [28]
In this circuit (Error! Reference source not found.), we have an
impedance Z0 connected in series. There are two Z0 impedances in the
circuit, and they share the same value. We also have a 2-volt voltage
source connected to the circuit. Let's calculate the voltage across Z0.
Calculating the current (I) flowing through the circuit: According to
Ohm's law:
I = V / Z_total
I = 2 volts / (2 * Z0)
I = 1 / Z0 Amperes
If Z_naught is equal to 1 ohm, then the current (I) would be 1 Ampere.
Using the voltage division rule to find the voltage (V) across Z0:
V = Z0 / (Z0 + Z0) * 2 volts
V = (Z0 / 2Z_naught) * 2 volts
V = 1 volt
So, the voltage drop across Z0 is 1 volt.
Applying Kirchhoff's voltage law (KVL): 2 volts - I Z0 – I Z0 = 0
By solving this equation, you can determine the values of voltages and
currents in the circuit.
Example:
So, let us challenge you through another circuit to calculate the voltage
V out as shown in Figure 3‑19. Imagine you are given a supply voltage
Vin (or input voltage Vin), and there are three resistors: R1, R2, and R3,
connected as shown. We see that R2 and R3 are in parallel, and to this
parallel combination of R2 and R3, we have R1 in series with that. The
total resistance, RT, can be written as R1 + R2 || R3 (i.e. R2 in parallel
with R3).
Now, Vout can be calculated using the voltage division rule. Vout can
also be calculated by applying Ohm's law. You calculate the total current
flowing into the circuit, and then you calculate the voltage drops across
each resistor.
Figure 3‑19 currents and voltage calculations [29]
Here's how you can calculate the current (I) that flows through the
circuit: I = Vin / RT, where RT is R1 + R2 || R3 (i.e. R2 in parallel with
R3).
Now, let's calculate the individual voltage drops:
1. Voltage across R1 is VR1 = I * R1
2. Voltage across R2 is VR2 = IR2
3. Voltage across R3 is VR3 = IR3
These currents get divided into two parts, IR2 and IR3, which, when
multiplied by their respective resistors, provide the voltage drops across
the individual resistors. Now, for Vout, we need to find the voltage
across the parallel combination of R2 and R3, which we can call RX:
RX = (R2 * R3) / (R2 + R3)
Now, Vout can be calculated as follows:
Vout = (RX / (RX + R1)) * Vin
This equation helps you calculate Vout. Using the voltage division rule,
you can simplify it by noting that Vout is equal to the voltage across RX,
which is the voltage across R2 and R3. So, Vout = VR2 = VR3.
So, in summary, you can calculate V out using the voltage division rule
or Ohm's law.
Applying current division in
circuits
Figure 3‑20 Schematic of an electrical circuit illustrating current
division. Notation RT refers to the total resistance of the circuit to the
right of resistor RX [30]
In this circuit, we will explore the application of the current division
rule. The circuit consists of a resistive network with a current source
labeled IT, which represents the total current entering the circuit. The
direction of the arrow indicates the current's flow. We have three
resistors: R1, R2, and R3, as well as an additional resistor labeled RX,
through which the current IX flows. The total equivalent resistance of R1,
R2, and R3 is represented as RT.
If we are tasked with calculating the current flowing through RX (IX), we
can apply the current division rule. The formula for calculating IX is as
follows:
IX = (RT / (RT + RX)) * IT.
Here, RT represents the equivalent resistance of the three resistors (R1,
R2, and R3) connected in parallel. We divide this value by the sum of the
total resistance and RX and then multiply it by the total current, IT.
Suppose we need to calculate the current flowing through R3. In that
case, we follow a similar approach by first finding the equivalent
resistance of RX, R1, and R2, denoted as "RTA." The formula for
calculating the current through R3 (IR3) is:
IR3 = (RT / (R3 + RT)) * IT.
To calculate the current through R2, we find the equivalent resistance RTB
of RX, R1, and R3, using a parallel combination. Then, the formula to
determine the current through R2 (IR2) is:
R2 = (RT / (R2 + RT)) * IT.
This approach demonstrates how to use the current division rule to
calculate branch currents in any given circuit.
Chapter 4 Semiconductors
What is Semiconductor?
In this module, we aim to understand the concept of semiconductors. A
semiconductor is a material with electrical conductivity lying between
that of a conductor and an insulator (Figure 4‑1). Conductors, such as
copper and iron, exhibit high electrical conductivity, while insulators like
glass and rubber have low conductivity.
Figure 4‑1 Electrons and energy band in materials
Semiconductors, like gallium arsenide, germanium, and silicon, have
intermediate conductivity and are crucial for controlling and managing
electric current flow in electronic devices and equipment. To
comprehend how semiconductors function, let's examine their energy
band diagram. The energy band diagram illustrates the relationship
between the conduction band and the valence band in different materials.
In a conductor, the conduction and valence bands overlap, enabling
charge carriers (electrons or holes) to easily transition from the valence
band to the conduction band, facilitating electric current flow.
In contrast, insulators have a large energy gap between the valence and
conduction bands (greater than 5 electron volts), necessitating a
substantial energy input for charge carriers to move from the valence to
the conduction band. Consequently, insulators have limited electrical
conductivity.
Now, let's focus on semiconductors. Semiconductors possess a smaller
energy band gap compared to insulators, making it relatively easier for
charge carriers to transition between the valence and conduction bands.
This transition can occur through processes such as doping or
temperature changes. As a result, semiconductors offer the ability to
control and manage electric current flow, making them a fundamental
choice in electronic equipment, devices, circuits, and systems.
Let's take a closer look at the energy band diagram of a semiconductor,
Figure 4‑2. Semiconductors come in two primary types: p-type and n-
type semiconductors. When a pure semiconductor is doped with specific
impurities, it transforms into an extrinsic semiconductor. Extrinsic
semiconductors further divide into two categories: n-type and p-type
semiconductors.
In Figure 4‑2, we focus on the energy band diagram of an n-type
semiconductor. Like any semiconductor, it consists of key energy bands:
the dark gray region represents the conduction band, while the white
dotted area symbolizes the valence band. However, in a doped
semiconductor, such as an n-type semiconductor, there's an additional
energy band known as the donor band, indicated by the dashed lines.
This donor band signifies the presence of excess electrons, making it an
n-type semiconductor, where a surplus of electrons is available for
conduction.
Energy band in Semiconductors
Figure 4‑2 Energy band in semiconductors [31]
From an energy level perspective, we have the following designations:
EC (Conduction Band Energy Level)
EV (Valence Band Energy Level)
EF (Fermi Level)
The Fermi level (EF) plays a pivotal role in semiconductor conductivity.
It represents the energy level at which an electron must attain sufficient
energy to transition from the valence band to the conduction band,
enabling the flow of electric current. In n-type semiconductors, the
presence of the donor band ensures a surplus of electrons, making it
easier for current to flow.
Doping a semiconductor, as depicted in this energy band diagram,
signifies that the material has been altered to become an n-type
semiconductor, characterized by a substantial number of charge carriers,
in this case, electrons, available to facilitate efficient electrical
conduction within the material
Doping in Semiconductor
Doping involves the introduction of impurities into a pure
semiconductor material like silicon. This process enhances or increases
the semiconductor's conductivity. Let's delve into the details:
In its pure form, silicon atoms each possess four valence electrons in
their outer orbits. For instance, every silicon atom has four electrons in
its outer shell, and they form bonds with neighboring silicon atoms. This
bonding is known as covalent bonding, where one electron from one
silicon atom pairs with one electron from a neighboring silicon atom,
creating a bond. This covalent bonding occurs throughout the crystal
lattice structure.
n-type p-type
Figure 4‑3 Silicon doping transforming the pure form into n-type and p-
type semiconductors
Now, when we introduce an impurity into the pure silicon, let's say a
semiconductor material like Antimony (Sb), which is termed a 'donor,'
something interesting happens. Antimony has five electrons in its outer
shell, as indicated by the five electrons in Figure 4‑3. Four of these
electrons form covalent bonds with neighboring silicon atoms, just like
in the intrinsic silicon crystal. However, there's an extra or fifth electron
left unbound.
This introduction of an impurity is referred to as 'doping.' Due to this
doping process, we now have an excess of electrons within the crystal
lattice. Consequently, the intrinsic silicon, which was initially a pure
semiconductor, transforms into an 'N-type' semiconductor. The 'N' stands
for 'negative' because of the excess electrons introduced by the donor
impurity.
Mathematically, the excess electron density in the N-type semiconductor
can be described as the difference between the number of electrons
introduced by the donor impurity (such as phosphorus) and the number
of holes (missing electrons) in the intrinsic silicon crystal lattice. This
excess electron density, represented by 'n,' plays a crucial role in the
conductivity of the N-type semiconductor.
Let's examine another example to understand the process of doping in
silicon material. In the second image of Figure 4‑3, we can observe
silicon atoms, each of which has four valence electrons in their outer
orbits. These electrons are referred to as silicon valence electrons. It's
important to note that each silicon atom possesses four such outer
electrons, establishing the foundation for covalent bonding.
The covalent bonding is illustrated in the diagram, where an electron in
the outer shell of each silicon atom forms a bond with a neighboring
electron present in the outer shell of an adjacent silicon atom. This
bonding pattern is a characteristic feature of pure silicon, which is a
pristine semiconductor.
Now, let's introduce an external impurity into this pure silicon material.
For this example, we'll consider Boron (B) as the impurity. Boron, unlike
silicon, has only three valence electrons in its outer shell. When we add
a boron atom (depicted in the image), it contributes its three valence
electrons to form covalent bonds with the valence electrons of
neighboring silicon atoms. However, there is a deficiency or void of one
electron because boron has only three electrons to offer.
This deficiency or absence of an electron creates what is known as a
'hole' within the crystal lattice structure. This hole is essentially a vacant
position where an electron should be. As we introduce more boron
atoms, more such holes are created in the silicon structure due to the
shortage of electrons from the boron impurity.
This process of introducing an impurity, like boron, into the silicon
lattice creates numerous holes within the structure. As a result, the
material becomes a 'p-type' semiconductor. The 'p' in 'p-type' signifies
the prevalence of positively charged holes in the material due to the
absence of electrons in these positions.
In summary, the doping process involving boron impurity transforms the
pure silicon, which initially had a balanced number of electrons and
holes, into a p-type semiconductor characterized by a surplus of
positively charged holes.
Doping plays a crucial role in the fabrication of semiconductor devices,
as it allows precise control over the electrical behavior of these
materials. By strategically introducing impurities, engineers can create
N-type and P-type regions within a semiconductor device, enabling the
development of diodes, transistors, and other electronic components that
form the basis of modern electronics. This control over electrical
properties is essential for the design and operation of semiconductor
devices.
Here are some of the most commonly used n-donors and p-acceptors:
N-Dopants (Electron Donors):
1. Phosphorus (P): Phosphorus is one of the most commonly
used n-dopants. It has five valence electrons, and when it is
introduced into silicon, for example, it contributes an extra
electron, creating an excess of electrons in the crystal lattice.
2. Arsenic (As): Arsenic, like phosphorus, has five valence
electrons. When doped into silicon, it also acts as an n-dopant
by providing an additional electron.
3. Antimony (Sb): Antimony is another element with five
valence electrons, making it suitable as an n-dopant in
semiconductors. It donates an extra electron to the crystal
structure.
P-Dopants (Hole Acceptors):
1. Boron (B): Boron is one of the most commonly used p-
dopants. It has only three valence electrons, and when
incorporated into a semiconductor lattice, it creates 'holes' by
forming bonds with neighboring atoms, leaving vacancies
where electrons should be.
2. Gallium (Ga): Gallium is another element used as a p-dopant.
It also has three valence electrons and introduces 'holes' into
the crystal structure.
3. Indium (In): Indium, like boron and gallium, has three
valence electrons and can be used to create p-type
semiconductors by introducing holes into the material.
PN Junction
Figure 4‑4 PN junction and diode symbol [32]
In Figure 4‑4, we observe the concept of P-N junction formation. When
you combine a p-type semiconductor material with an n-type
semiconductor material, you create what is known as a PN Junction.
This fundamental junction forms the basis of various semiconductor
devices, including diodes and transistors.
A PN junction is a crucial component in semiconductor electronics. It's
formed by joining a region of p-type material, where there's an excess of
'holes' (positive charge carriers due to missing electrons), with a region
of n-type material, characterized by an excess of electrons. At the
boundary where these two types of materials meet, interesting electronic
properties emerge.
This junction allows for the controlled flow of charge carriers,
depending on the voltage applied. In a forward bias, it facilitates the
flow of current, while in a reverse bias, it acts as a barrier to current
flow. This property of the PN junction is harnessed in diodes to allow
current flow in one direction only, making them useful for rectification
and signal switching. Transistors, on the other hand, leverage the PN
junction's ability to amplify and control current for various electronic
applications. This junction plays a pivotal role in various semiconductor
devices due to its unique electronic properties.
PN junction semiconductor diodes
Referring to Figure 4‑4, we are looking at two types of silicon materials:
p-type silicon and n-type silicon. Both of these materials belong to the
category of semiconductors. P-type semiconductor is created by
introducing an acceptor-type impurity, while n-type silicon is formed by
adding a donor-type impurity to the silicon material. Both p-type and n-
type silicon are examples of extrinsic semiconductors, achieved through
a process called doping.
Now, at one end of the p-type silicon, we have a terminal referred to as
the 'anode,' as indicated here. At the other end of the n-type silicon
semiconductor, we have a terminal known as the 'cathode.' When these
two types of semiconductors are joined together, the region where they
meet is called the 'junction.' This combination gives rise to what we call
a 'PN junction.' The entire device is commonly known as a 'PN Junction
Diode.'
In Figure 4‑4, you can see the symbol representing a PN junction diode,
where one terminal is designated as the 'anode,' and the other terminal is
labeled as the 'cathode.' It's essential to remember this symbol when
working with PN junction diodes, as it is widely used in electronic
schematics and diagrams.
Physics of PN junction
semiconductor devices
In this module, we aim to understand the functioning of a PN junction
(Figure 4‑5). Here, you can see a P-type semiconductor referred to as P
(with an abundance of holes) and an N-type semiconductor (with a
surplus of electrons). P-type semiconductors predominantly contain
holes as their majority charge carriers, while N-type semiconductors are
rich in electrons as the majority charge carriers.
Figure 4‑5 Physics of PN junction semiconductor devices
When a P-type semiconductor is joined with an N-type semiconductor,
as indicated by the line here, electrons (majority carriers) and holes
(positive charge carriers, majority in P-type) attempt to diffuse from one
region to another due to electrostatic forces. This phenomenon is known
as the diffusion process. Electrons from the N-type region migrate to the
P-type region and fill some of the holes (shown in blue). Conversely,
some holes travel from the P-type region to the N-type region and fill
with electrons (shown in blue). This intermingling of charge carriers
occurs within the space charge region or depletion region.
The depletion region forms a built-in potential or built-in voltage across
the PN junction, denoted as ΔV. This voltage arises due to the presence
of positive ions (red symbols) and negative ions (blue symbols) within
this region. It's important to note that the depletion region lacks free
electrons or holes; instead, it contains these ions.
Next, let's discuss some electrical quantities concerning the charge
distribution along the x-axis. In the depletion region (the blue zone with
negative ions), the charge reaches its maximum. At the junction, the
charge is zero, and it increases negatively in the depletion region where
negative ions are present. Conversely, in the red region where positive
ions exist, the charge increases positively. This charge distribution
defines the electric field.
The voltage profile shown here demonstrates that there is a voltage in
the depletion region (ΔV), which is the built-in voltage. To facilitate the
flow of electricity in this PN junction, an external voltage is applied
across it. By means of this applied electric field, we need to overcome
the depletion region voltage, which is the barrier voltage or built-in
voltage. This allows electrons (or holes) to traverse from one region to
another, effectively completing the circuit.
In essence, when two semiconductor materials are joined to form a PN
junction, the application of an external voltage enables the flow of
current. By overcoming the internal barrier voltage, the electric field
facilitates the movement of electrons (or holes) from one region to
another, resulting in conduction.
Chapter 5 Semiconductor Devices
Diode
PN junction silicon and germanium
diode
Silicon diode [33] Germanium diode [34]
Figure 5‑1 real diodes
Figure 5‑1 shows real PN Junction silicon and germanium diodes, with
identifiable terminals. The end with the silver ring (silicon) or black ring
(germanium) is referred to as the cathode, while the opposite terminal is
called the anode. It's important to remember that the symbol for this diode
is as shown in Figure 5‑1.
Silicon diodes typically have a cut-in voltage of around 0.7 volts for
standard silicon diodes. This means that they begin to conduct current when
a forward voltage of approximately 0.7V is applied across them.
Germanium diodes have a lower cut-in voltage compared to silicon diodes.
They typically have a cut-in voltage of around 0.3 to 0.4 volts. This makes
them suitable for low-voltage applications. Silicon diodes are widely used
in various electronic circuits, including rectifiers, voltage clippers, and
signal clamps. Germanium diodes have been largely replaced by silicon
diodes in modern electronics due to their lower performance at higher
temperatures.
PN junction diode IV current-
voltage curve
Figure 5‑2 PN junction diode input-output characteristics [35]
In this segment, we observe the typical IV (current-voltage) characteristics
of a PN Junction diode (Figure 5‑2). The IV characteristics depict the
relationship between current (I) and voltage (V). There are two common
semiconductor diodes: one made of germanium and the other of silicon.
The red curve represents the IV curve for the silicon diode, while the blue
curve represents the IV curve for the germanium diode.
When you forward bias a PN Junction diode, it means you apply a positive
voltage to the P-type material and a negative voltage to the N-type material.
This action moves the diode into forward bias, as illustrated. The cut-in
voltage, also known as the threshold voltage, is approximately 0.5 volts for
silicon diodes. At this point, the diode begins to conduct and acts as a short
circuit. Current flows from P to N, and the circuit is closed. Until the
voltage reaches 0.5 volts on the x-axis, the current in the diode remains
minimal. The same behavior applies to germanium diodes, but their cut-in
voltage is approximately 0.2 volts. Beyond this voltage, the current starts to
increase.
Now, let's discuss the reverse bias characteristics. When the diode is
reverse-biased, as indicated in the diagram, the current through the diode
remains very small as the voltage is increased. This is due to the fact that
the junction is reverse-biased and effectively acts as an open switch.
However, as you continue to increase the voltage, there is a sudden
breakdown, and heavy current begins to flow through the diode. This is
known as the breakdown region, and it is experienced by both silicon and
germanium diodes.
Understanding these IV characteristics is crucial for using diodes
effectively in electronic circuits.
Region of operations of pn junction
semiconductor diode | I-V curve
Figure 5‑3 Diode regions of operation [36]
1. Forward Bias Region: When a positive voltage is applied to the
diode (forward bias), starting from zero voltage and gradually
increasing, there is initially a small current flow. However, as the
applied voltage reaches the diode's threshold voltage, often
referred to as the "cut-in" voltage (VD), there is a sudden
exponential increase in the diode's current. This region,
represented by the green curve, signifies the forward biasing of
the diode. In this region, the diode is conducting current
efficiently.
2. Reverse Bias Region: Conversely, when a negative voltage is
applied across the PN Junction diode (reverse bias), starting from
zero and increasing in the negative direction, the diode exhibits
minimal or ideally zero current flow. This region, shown by the
blue curve, indicates reverse bias operation. In this situation, only
a negligible amount of current, known as reverse leakage current,
flows through the diode.
3. Breakdown Region: In the breakdown region, as the reverse
voltage continues to increase, there comes a point where the
diode suddenly exhibits a significant increase in current. This
phenomenon is known as breakdown. When the reverse voltage
reaches a certain critical value, known as the breakdown voltage,
or reverse breakdown voltage (VBR), the diode experiences an
avalanche effect, and the current surges rapidly. Operating the
diode in this region can lead to damage.
These three regions are the fundamental operating characteristics of a PN
Junction diode. Understanding and properly utilizing these characteristics
are essential for designing and using electronic circuits involving diodes.
Bipolar junction transistor
Figure 5‑4 How BJT looks like [37]
In Figure 5‑4, we observe a transistor, specifically a BJT (Bipolar Junction
Transistor). A BJT is a semiconductor device, and it derives its name from
the fact that it conducts electricity using two types of charge carriers:
electrons and holes. Depending on the type of charge carriers it utilizes for
conduction, it can be classified as either an NPN or PNP transistor.
There are two main types of transistors:
1. NPN Transistor: This type of transistor employs electrons to
carry the current.
2. PNP Transistor: PNP transistors use holes to conduct the current.
Transistors typically feature three terminals:
1. Emitter
2. Base
3. Collector
These components are found in various forms, including integrated circuits
(ICs), metal cans, and ceramic packages. That concludes our overview of
transistors. We have discussed transistors in detail in the following sections.
Transistor types and configurations
Bipolar junction transistor (BJT)
operation
Figure 5‑5 Bipolar junction transistor operation and biasing [38]
The diagram in Figure 5‑5 illustrates the operation of an NPN transistor. In
this configuration, 'N' represents the N-type semiconductor, while 'P'
represents the P-type semiconductor. The transistor consists of three layers,
with a P-layer sandwiched between two N-type layers, hence the term 'NPN
transistor.'
Let's examine the voltage sources connected to the transistor. The first
voltage source, denoted as 'VBE,' is connected between the base and
emitter. The positive terminal of VBE is linked to the base, while the
negative terminal connects to the emitter. This arrangement forward-biases
the PN junction, with the positive side connected to the P-type
semiconductor and the negative side to the N-type semiconductor.
Now, let's focus on the other voltage source, 'VCB,' connected across the
collector and base of the transistor. The positive terminal is linked to the
collector, while the negative terminal connects to the base. This
configuration reverse-biases the junction. Here, the positive side connects
to the N-type material, and the negative side connects to the P-type
semiconductor.
This combination results in one junction being forward-biased, while the
other is reverse-biased. This is the fundamental condition for a transistor to
operate in its active region.
Turning our attention to the circuit, we have the collector current, denoted
as IC, flowing into the collector. Simultaneously, the emitter current, IE,
flows out. Additionally, there is a base current, IB, flowing into the base
region. Mathematically, IE is the sum of IB and IC.
From a physics perspective, the direction of electron flow is opposite to
conventional current flow. Electrons move from the emitter towards the
collector. Some of these electrons recombine with the holes in the base
region, signifying recombination. This figure effectively explains the
operation of an NPN transistor when biased in its active region.
BJT and its symbols
An NPN transistor consists of three semiconductor layers: N-type, P-type,
and N-type, arranged in that order. Referring to Figure 5‑6, the arrow in the
symbol represents the direction of conventional current flow, which goes
from the collector to the emitter, with little to no current flowing into the
base in an ideal transistor. Electrons, which are negatively charged, flow in
the opposite direction, from emitter to collector, to allow this conventional
current flow.
Figure 5‑6 BJT symbols [39]
In case of a PNP transistor, P stands for P-type semiconductor and N stands
for N-type semiconductor. In this configuration, a P-type semiconductor
material is sandwiched between two adjacent N-type semiconductor
materials. Referring to Figure 5‑6, the arrow in the symbol still represents the
direction of conventional current flow, which goes from the emitter to the
collector in a PNP transistor, with little to no current flowing into the base in
an ideal transistor. Electrons, which are negatively charged, flow in the
opposite direction, from collector to emitter, to allow this conventional
current flow.
Transistor as a switch and amplifier
Figure 5‑7 Transistor as a switch/amplifier [40]
Let's delve into the operation of a transistor through this simple circuit
(Figure 5‑7). As you can see, it's an NPN transistor, characterized by the
direction of current flow. In this configuration, the N-region is sandwiched
between two P-regions, and the current flows from the collector to the
emitter, hence the name NPN transistor. Now, when an input signal is
applied to the base of the transistor, the base-emitter junction becomes
forward-biased. This results in a small current called the base current (IB)
flowing into the input side of the circuit. The transistor's gain, denoted by β
(beta), amplifies this base current to produce the collector current (IC) at
the output side of the circuit. In essence, the small input signal is
transformed into a larger output signal, illustrating the transistor's
amplification capability.
This same circuit, aside from acting as an amplifier, can also function as a
switch. Imagine when the input signal is at zero volts; this effectively turns
off the input side of the circuit, causing no input current, and consequently,
no output current. Therefore, the output voltage (Vout) is tied to VCC.
Conversely, when a high voltage signal is applied, it forward-biases the
base-emitter junction, allowing a base current (IB) to flow, which, when
amplified by the transistor, results in an output current (IC). This current
passing through the resistor (R) causes a voltage drop, so the output voltage
(Vout) becomes VCC - IC * R, which is lower than VCC. This behavior
effectively demonstrates how the transistor can function as a switch.
Example
Figure 5‑8 Transistor as a switch
In this circuit (Figure 5‑8), we have an NPN transistor with distinct
collector, base, and emitter terminals. The emitter is grounded, the collector
is linked to a 6-volt supply, and the base is under the control of a switch.
When the switch is in the open position, it disrupts the current flow through
the resistor since the circuit is incomplete. Consequently, no current enters
the input side of the transistor. This leads to the transistor being in an "off"
state because there is no current on the input side. Therefore, the transistor
acts as an open circuit, and the light bulb connected to it remains unlit.
However, when the switch is closed, it closes the circuit, enabling a current
(IB) to flow through the resistor and into the input side of the transistor. We
can apply Kirchhoff's voltage law to the input side of the transistor, which
can be expressed as follows:
6 V - IB * 1 kΩ - VBE = 0.
Here, VBE represents the forward voltage drop across the base-emitter
junction of the transistor, typically around 0.6 to 0.7 volts for a silicon
transistor. As a result, the voltage across the input side decreases.
Considering the forward voltage drop, IB induces a corresponding current,
known as the collector current (IC), to flow from the collector to the emitter
on the output side of the transistor. This establishes a conduction path from
the supply voltage to ground, and the light bulb connected in this path will
illuminate.
In summary, when the switch is closed, the transistor enters the "on" state,
permitting current to flow from the collector to the emitter, and the bulb
lights up. Conversely, when the switch is open, the transistor is in the "off"
state, interrupting the current flow and keeping the bulb off. This illustrates
how a transistor functions as a switch, regulating current flow in an
electronic circuit.
Field effect transistor (FET)
How does a Field Effect Transistor
works? | FET vs BJT
This current is controlled by an input voltage applied at the gate. This is
why it's called a Field Effect Transistor, as the applied electric field at the
gate controls the current in this device.
Now, let's examine the structure of this semiconductor device. It consists of
a P-type semiconductor material as shown, known as the body of the FET,
with two N+ doped regions within it. The N+ regions have plenty of
electrons, while the P-type material has plenty of holes. The P-type
material contains holes as the majority carriers and electrons as minority
carriers, whereas the N-type material contains electrons as the majority
carriers and holes as minority
(a) (b)
Figure 5‑9 Field effect transistor [41]
carriers. In total, there are four pins for the device: body (a P-type
substrate), gate, source, and drain.
On the body of the P-type semiconductor, an insulator material is
deposited, known as gate oxide. Typically, this material is SiO2 (silicon
dioxide), although silicon nitride can also be used. On top of the insulator,
there is a layer of metal or polysilicon, which is why it's also called a
Metal-Oxide-Semiconductor or MOS. In specific cases, it's referred to as a
MOSFET (Metal-Oxide-Semiconductor Field Effect Transistor).
Now, there are two types of FETs: MOSFET and Junction Field Effect
Transistor (JFET). JFET is an example of a unipolar device, unlike the
Bipolar Junction Transistor (BJT), which is bipolar. In a BJT, current
conduction involves both types of charge carriers, electrons, and holes. In
contrast, in FET, current conduction occurs due to only one type of charge
carrier at a time, either electrons or holes.
Referring to the (b) part of (Figure 5‑9), when you apply a voltage at the
gate (for example, a positive voltage), positive charge carriers accumulate
on top of the metal layer. This accumulation causes negative charge carriers
(minority carriers) in the body to move closer to the oxide-semiconductor
interface. Once a sufficient positive voltage is applied to overcome the
threshold voltage of the transistor, enough negative charge carriers
accumulate at the oxide-semiconductor interface, allowing an electric field
to form across the drain and source regions. This results in current flowing
from the drain to the source, completing the circuit. Removing the gate
voltage stops current conduction, and reapplying it allows current to flow
again, provided there is a voltage across the drain and source.
In summary, the voltage applied at the gate controls the current in the
channel region, which is called the channel of the FET. That's the basic
concept of Field Effect Transistors.
How Field effect transistor controls
current | FET full operation
Part a Part b
Part c
Figure 5‑10 Field effect transistor
Let's analyze Field Effect Transistors (FETs) (Refer Figure 5‑10, part a).
We've seen that when we apply a positive voltage relative to the body
(which is tied to the negative terminal) at the gate, it deposits positive
charges on top of the gate, which is on top of the metal. If this positive
voltage at the gate is sufficient, it will attract electrons from the P-type
body towards the gate oxide and the P-type semiconductor interface.
As a result, an electric field is created, and an electric force is applied from
the positive charges to the negative charges, following Coulomb's law.
Now, if we apply a positive voltage to the drain pin of the transistor while
tying the source to the ground or the negative, (part a), remember that the
P-type body and the source are at the same negative potential, this situation
floods the N-type semiconductor, and the N+ regions at both ends of the
transistor, with electrons. In the channel, many electrons accumulate,
creating a complete path from source to drain. Thus, electrons flow from
drain to source, and this current is called IDS (drain-to-source current).
Now, let's consider diode conditions (refer to part b in the figure). A P-type
semiconductor is connected to the N-type semiconductor, with the N-side
tied to the positive and the P-side to the negative. This diode is reverse-
biased. Similarly, another diode’s terminals are also connected to the
negative voltages at both the ends. This results in a completely reverse-
biased diode. Consequently, the junction diodes do not become forward-
biased, which is a favorable situation. Forward-biased diodes in this
configuration could interfere with the operation of the FET.
Moving on to the third situation (part c), let's say we apply a negative
potential to the gate, with the body tied to the positive potential, creating
opposite polarity conditions. This situation causes many negative charge
carriers to accumulate on top of the gate, attracting many positive charge
carriers to the gate oxide and the P-type semiconductor interface. This
creates an electric field, with the direction of electric field lines and
electrostatic force (direction of arrows) as shown.
Now, will current flow when you apply a voltage difference between source
and drain?
In this case, the N+ regions on either end of the transistor have many
electrons, but in the middle, the channel has many holes, which are positive
charge carriers. Thus, there is no current flow from source to drain or drain
to source.
Regarding the diodes, the P-body is tied to the positive terminal and the N-
side to the negative, resulting in a forward-biased condition for this diode.
Similarly, for another diode, both P-body and N+ region are connected to
the positive voltages. This diode is not forward-biased. Essentially, for part
c, no current will flow through the transistor even if you reverse the
polarity of source and drain and only one of the junctions are forward-
biased.
In summary, we've examined four conditions: gate positive relative to the
body, gate negative relative to the body, drain positive relative to source,
and source positive relative to drain. These biasing conditions help us
understand the current flow and switching action of the FET.
Depletion region | n-channel FET |
circuit
Figure 5‑11 How voltage at the Gate controls the depletion region width in
n-channel FET [42]
In the provided circuit description, we have a Field Effect Transistor (FET)
configuration (Figure 5‑11).
To the left, there's an N-type silicon semiconductor bar, and within this N-
type material, there are two P-regions, resulting in the formation of two P-
N junctions.
On either side of these P-regions, there are two gate pins, which are shorted
together and considered as a common gate. Additionally, there are two
more pins labeled as source and drain, totaling three pins for this FET.
A voltage source is applied across the drain and source, with the positive
terminal connected to the drain and the negative terminal connected to the
source. In this setup, the N-type semiconductor, represented by the blue
region, acts as a conductive path through which the voltage source supplies
current. The current flows from drain to source, as indicated by the arrow's
direction. It's important to note that due to the presence of two P-N
junctions, depletion regions are formed within the FET's channel, which
corresponds to the blue region referred to as the channel of the transistor.
This configuration represents an N-channel FET because current flows
through the N-type material.
Now, let's examine the circuit on the right. Here, a battery is connected to
the gate terminal with respect to the source terminal. The negative terminal
of the battery is connected to the gate of the transistor, while the positive
terminal of the battery is connected to the source.
In this setup, a negative voltage is applied to the P-type semiconductor on
one side (gate-source region), and a positive voltage is applied to the drain
on the other side. Consequently, the P-N junction between the drain and the
gate becomes reverse-biased, resulting in the depletion region indicated in
the diagram.
Similarly, the source terminal of the N-type material is connected to the
negative potential, and the gate terminal is connected to the negative
potential as well, or alternatively, the source is connected to the positive
potential, and the gate is connected to the negative potential. This setup
also results in a reverse-biased P-N junction and the corresponding
depletion region.
The magnitude of the negative voltage applied to the gate determines the
size of the depletion region. As the negative voltage increases, the
depletion region expands, reducing the current flowing from drain to
source. Conversely, applying a positive voltage to the gate with respect to
the source narrows the depletion region, allowing a larger current to flow
from drain to source.
In summary, the gate-source voltage controls the current flowing through
the FET's channel. By adjusting this voltage, the current can be effectively
controlled. Figure 5‑12 shows JFET symbols for both the n-channel and p-
channel.
Figure 5‑12 JFET symbols
Metal Oxide Semiconductor Field effect transistor (MOSFET)
Figure 5‑13 MOSFET, showing gate (G), body (B), source (S), and drain
(D) terminals. The gate is separated from the body by an insulating layer
(light pink) [43].
Let's dive into the details of the basics of MOSFETs.
What is a MOSFET? MOSFET stands for Metal Oxide Semiconductor
Field Effect Transistor. In Figure 5‑13, you can see the components of a
MOSFET. It consists of a silicon substrate, also referred to as the body of
the transistor or device. On top of this substrate, there are two terminals:
one is called the drain, and the other is the source. Separating these
terminals is an insulating layer, typically silicon dioxide or silicon nitride,
and above this insulating layer, there is a metal or polysilicon layer. This
combination of metal and oxide constitutes the semiconductor. MOSFETs
are known for their switching capabilities. For a comprehensive
understanding of their operation, we highly recommend checking out our
course on CMOS VLSI titled "CMOS VLSI by Dr. Vinayak Pachkawade".
Figure 5‑14 FET symbols [44]
MOSFET Symbol The MOSFET symbol represents an extension or
modification of the Junction Field Effect Transistor (JFET). MOSFETs
operate by applying an electric field, hence the term "Field Effect
Transistor." There are two types of MOSFETs: P-channel and N-channel
(Figure 5‑14). The symbols for these are as follows:
P-Channel MOSFET: The arrow points outward.
N-Channel MOSFET: The arrow points inward.
Each MOSFET type has three terminals: Drain (D), Gate (G), and Source
(S). Additionally, there is a fourth terminal called the Body (B).
How does MOSFET transistor
look?
Figure 5‑15 MOSFET Transistors [45]
We see a transistor known as a MOSFET, which stands for Metal-Oxide-
Semiconductor Field-Effect Transistor Figure 5‑15. This is a three-terminal
device and falls under the category of discrete electronics semiconductor
devices. As you can observe, there are various variations and versions of
these transistors. A MOSFET comprises three terminals: the gate, the drain,
and the source. To understand which terminal serves what purpose, you can
refer to the datasheet specific to the particular transistor. You can locate the
code on the body of the transistor, often under a microscope, and enter it
into Google to retrieve the relevant datasheet, providing comprehensive
information about that MOSFET. Typically, MOSFETs can be classified as
N-channel or P-channel, and they come in two primary types:
enhancement-type MOSFETs, which are typically off transistors, and
depletion-type MOSFETs, which are usually on transistors. In the case of
enhancement-type transistors, applying a gate voltage causes an increase in
current. This effectively changes the transistor's state from off to on. On the
other hand, for depletion-type transistors, the transistor is normally in the
on state. By applying a voltage to the gate, you can gradually turn it off by
reducing the current level. This provides a fundamental understanding of
MOSFETs.
Understanding MOSFETs
(a)
(b)
Figure 5‑16 Scheme of metal oxide semiconductor field-effect transistor (a)
[46], Cross section of two transistors in a CMOS gate, in an N-well CMOS
process (b) [47]
Let's dive into the world of MOSFETs to gain a deeper understanding.
MOSFETs are majority carrier devices, which means that the current
flowing through them is primarily due to either electrons or holes. To
comprehend their operation, let's examine the structure of a MOSFET.
In Figure 5‑16 (a), we have a P-type semiconductor, serving as the body or
substrate of the device. Inside this substrate, we find two heavily doped
regions labeled as N+ and N+, forming the source and drain, respectively.
Separating these regions is an insulating layer, typically made of silicon
dioxide or silicon nitride. On top of this insulating layer, we have a metal or
polysilicon layer, which acts as the gate.
The MOSFET, with its distinct structure, is capable of controlling current
flow from drain to source when a voltage is applied to the gate terminal. It's
important to note that MOSFETs come in two fundamental types: N-
channel and P-channel.
N-Channel MOSFET (NMOS): In an N-channel MOSFET (NMOS), the
majority carriers responsible for current conduction are electrons. When a
positive voltage is applied to the gate, a sufficient number of electrons
gathers in the channel between the source and the drain. With the drain at a
positive terminal and the source at a negative terminal, current flows from
source to drain. In other words, electrons move from the source to the
drain. The voltage applied at the gate controls this current.
P-Channel MOSFET (PMOS): Conversely, in a P-channel MOSFET
(PMOS), the majority carriers in the channel are holes. To create a PMOS
transistor, a P substrate is used, with two P+ regions forming the source and
drain. Like the NMOS, there is an insulating layer and a gate terminal on
top. When an appropriate voltage is applied to the gate, holes in the channel
between source and drain enable current flow from drain to source. This
current, too, is controlled by the gate voltage.
In MOSFETs, N-channel or P-channel, operate as majority carrier devices.
NMOS relies on electrons as majority carriers, while PMOS employs holes.
The voltage applied at the gate governs the flow of current in the
conducting channel between the source and the drain.
Understanding MOSFETs is essential, as these devices play a pivotal role
in modern electronics. Stay tuned for further insights into their operation
and applications.
Chapter 6 Diode and Transistor Circuits
Diode circuits
Diode half wave rectifier | AC to
DC conversion
Figure 6‑1 Diode half-wave rectifier - AC to DC converter [48]
Let's examine how this circuit operates (Figure 6‑1). In this
configuration, there is an alternating current (AC) voltage source
connected to the primary winding of the transformer. Across the
secondary winding of the transformer, we obtain a reduced AC voltage,
which is why this transformer is referred to as a step-down transformer.
The primary function of this transformer is to convert a high-magnitude
AC voltage on the primary side into a lower-magnitude AC voltage on
the secondary side.
The secondary voltage, present across points A and B, is sinusoidal in
nature. This means it exhibits both positive and negative half cycles.
During the positive half cycle of the alternating signal, the diode
becomes forward-biased and functions as a closed switch. Consequently,
the positive signal flows through the resistor, and this is reflected in the
output waveform, where the positive half cycle is evident.
However, during the negative half cycle of the AC voltage, the diode
becomes reverse-biased, effectively acting as an open circuit.
Consequently, the signal cannot pass through the diode, and as a result, it
does not reach the output load resistors. This omission of the negative
half cycle from the output waveform characterizes the operation of this
circuit.
When the positive cycle repeats, the diode once again becomes forward-
biased, allowing the signal to pass through the load resistance.
Consequently, this signal is reflected in the output waveform.
In summary, this circuit functions as a half-wave rectifier or an AC-to-
DC voltage converter. It converts a sinusoidal AC waveform into a
pulsating DC waveform where the positive half cycles are preserved, but
the negative half cycles are omitted. However, it's important to note that
this output is not a pure DC voltage, as it exhibits variations over time.
For a constant voltage with minimal fluctuations, additional components
or circuits are required.
Full wave rectifier | AC to DC
conversion | diode circuits
Figure 6‑2 Full-wave rectifier - center tap [49]
In this circuit diagram, we will explore how the circuit in Figure 6‑2
operates. You have an AC voltage source depicted in the figure. The
transformer shown here is known as a Center Tap Transformer because it
has a center tap in its secondary winding. As a result, the voltage
developed will have the following polarity: positive, negative, and
negative, positive.
During the positive half-cycle in the secondary winding, the upper diode
conducts, enabling the signal to pass as shown in this diagram. The
voltage developed across the load resistor becomes the output voltage.
Then, during the negative half-cycle, the lower diode conducts, allowing
the current or signal to flow through the load resistance R. The voltage
developed across R is again as shown in this figure. Consequently, the
output exhibits this type of voltage, which is referred to as full-wave
rectified voltage.
This circuit effectively converts the primary AC waveform, an
alternating current signal, into a pulsating DC waveform, as depicted in
the output diagram. Thus, it functions as a full-wave rectifier.
Full wave bridge rectifier circuit |
AC to DC voltage converter
Figure 6‑3 Full-wave bridge rectifier [50]
Alright, let's examine this new circuit. In this circuit, we have a
sinusoidal AC voltage source with a frequency of 50 Hertz and an
amplitude represented as V1. Typically, this is the AC main supply we
receive in our homes, typically rated at 230 volts and 50 Hertz,
delivering a sinusoidal AC power supply.
Moving on, we notice four diodes arranged in the fashion shown. Two
ends of this diode circuit are connected to a sinusoidal voltage source
through the provided connections. The other two points of this circuit
lead to an output voltage measurement setup, as depicted here, with
reference to ground. Additionally, a capacitor and a resistor are
connected, and the output voltage is measured across this resistor,
denoted as V0 with respect to ground.
Now, let's delve into how this circuit operates and its application. With
this sinusoidal waveform, we encounter both positive and negative
cycles of oscillation, with a frequency of 50 Hertz. During the positive
half-cycle, the diodes 1 and 3 becomes forward-biased, allowing the
signal to flow through it and complete the circuit. This positive half-
cycle voltage passes through the load resistor, labeled as R1, and is
eventually grounded, forming a complete path.
This same configuration continues during the negative half-cycle when
the polarity changes. This time, the diodes 2 and 4 becomes forward-
biased, while the others become reverse-biased, again allowing the
signal to pass through and complete the circuit. The output voltage is
developed across the load resistor, with the positive and negative
terminals as shown.
Now, let's look at the output waveform. The dotted black lines represent
the first and second positive half-cycles. As you can see, the positive
half-cycles are preserved, resulting in a pulsating DC voltage. However,
it's important to note that this isn't a constant DC signal, which is
typically required to be stable without any signal pulsations or variations
over time.
To achieve a more constant DC signal, we introduce a capacitor. When
the output voltage, due to the positive half-cycle, passes across the
resistor R1, it charges the capacitor. Similarly, during the negative half-
cycle, the voltage across resistor R1 again charges the capacitor,
maintaining the same polarity. The capacitor's role is to charge up to the
peak of the waveform and then discharge slightly before repeating the
process. As a result, the red waveform represents an improved DC
converted signal with a nearly constant output. This capacitor acts as a
filter, removing ripples or variations from the output waveform.
However, if you were to remove the capacitor, the black shaded dotted
lines represent the pulsating DC output, showing the significance of the
capacitor in smoothing the signal.
In conclusion, this circuit is known as a full-wave bridge rectifier circuit,
as it converts an AC waveform into a nearly constant DC waveform. To
achieve the desired output, it's essential to choose suitable values for the
resistor (R) and capacitor (C) to determine the charging and discharging
time constants of the capacitor.
Transistor circuits
In this section, we are going to explore an important BJT (Bipolar
Junction Transistor) configuration. There are three key BJT
configurations in which the BJT can be biased: the common emitter
mode, common base mode, and common collector mode.
Let's focus on the common emitter mode of the BJT. In this circuit
diagram, the input is applied to the base, the emitter is grounded, and the
output is taken from the collector. Thus, the emitter serves as a common
connection for both the input (base) and output (collector). The input is
applied across the base-emitter junction (VBE), while the output is
represented as VCB.
Transistor in common emitter
mode
(a) (b)
Figure 6‑4 BJT in CE mode, Basic NPN common-emitter circuit
(neglecting biasing details) (a), and Adding an emitter resistor decreases
gain, but increases linearity and stability (b) [51]
This configuration allows the BJT to function as both a voltage and
current amplifier. It typically offers a high current gain, often around 200,
enabling small input currents to result in significantly larger output
currents.
Additionally, it can produce voltage amplification, translating small input
voltage variations into substantial output voltage changes.
However, there are specific characteristics and challenges associated with
this common emitter configuration. First, it provides a medium input
resistance, making it moderately sensitive to changes in input impedance.
Second, it exhibits a high output resistance, which affects its ability to
drive low-impedance loads effectively. Lastly, when a sinusoidal voltage
is applied at the input, the output is 180 degrees out of phase with the
input signal, resulting in a phase shift.
To address some of the limitations of this circuit, a modification called
"emitter degeneration" can be introduced. This involves adding a small
resistor (re) from the emitter to ground. This modification alters the
transconductance (gm) of the circuit, which is now expressed as gm = IC /
VBE. The transconductance is a measure of how the output current varies
with changes in input voltage.
Emitter degeneration significantly reduces the dependence on intrinsic
transistor parameters, such as gm, and makes the circuit's gain primarily
determined by the ratio of resistors, namely RC and Re. This modification
improves gain stability and increases the input dynamic range, reducing
distortion for larger input signals.
In summary, the common emitter configuration with emitter
degeneration offers improved performance over the basic common
emitter circuit, as it reduces gain dependency on transistor characteristics
and mitigates issues related to gain stability and input dynamic range.
Transistor in common base mode
Figure 6‑5 BJT in CB mode [52]
Let's take a closer look at this small circuit as shown in Figure 6‑5. What
we have here is an NPN bipolar junction transistor (BJT). As we know, a
BJT has three terminals or pins: the collector, emitter, and base. It's
identifiable as an NPN transistor because the arrow on the emitter side
indicates the direction of conventional current flow, which goes from the
collector to the emitter through this transistor.
Now, in this circuit, we can observe that the base of the transistor is
connected to the ground, and the input signal Vin is applied to the emitter.
The output is taken from the collector. Thus, we can say that the input is
applied to the emitter with respect to the base, making the input known as
VBE (Voltage between Base and Emitter). The output is taken from the
collector with respect to the base, making the output VCB (Voltage between
Collector and Base).
Here, the base is common between the input and output, meaning it's
connected to both the emitter and the collector. This configuration is
referred to as the 'common base' configuration of the transistor.
It's worth noting that you can also use a PNP transistor in the common
base mode, depending on your circuit requirements. As a practice
exercise, you can try drawing the PNP circuit in the common base
configuration.
DC Biasing of the Transistor |
Common Base | BJT
Figure 6‑6 example circuit of BJT in CB mode [53]
Let's take a closer look at this circuit and try to understand its operation.
This circuit takes an AC input signal and produces an AC output signal.
It contains a bipolar junction transistor (BJT) configured as an NPN
transistor.
As indicated by the arrow on the emitter side, the conventional current
flows from the collector to the emitter, classifying this transistor as an
NPN type. The BJT has three terminals: collector, base, and emitter. The
collector is made of an N-type semiconductor, the emitter is also N-type,
and the base is P-type.
To operate this circuit as an amplifier, it needs to be biased with DC
voltage sources to set its operating point or DC bias point. Let's examine
the biasing and circuit components in detail:
1. AC Input: An AC input signal is applied to the circuit.
Capacitor C1 filters out any DC component, allowing only the
pure AC signal to pass.
2. Emitter Biasing: To bias the transistor, a voltage source labeled
VEE is connected with its positive terminal to the base and the
negative terminal to the emitter, through resistor Re. This
biasing configuration forward-biases the base-emitter junction,
creating a voltage VBE across it.
3. Output Biasing: Another voltage source labeled VCC is
connected with its negative terminal to the base and its
positive terminal to the collector through resistor RC. This
configuration reverse-biases the base-collector junction.
The transistor operates in an active region due to the forward-biased
base-emitter junction and reverse-biased base-collector junction. The
biasing conditions enable the transistor to function as a common-base
amplifier, with the base serving as a common terminal between the
collector and the emitter.
Capacitor C2 filters out any DC component from the output signal,
providing an AC output at the designated node. In summary, this circuit
utilizes an NPN BJT operating in common-base mode as an amplifier. It
is crucially biased to ensure proper functionality.
How do you bias the transistor for
operation? | common base | PNP
BJT example
Figure 6‑7 Bias the transistor in common base using PNP BJT example
[54]
Let's take a closer look at another circuit example. Here, we can see a
PNP transistor connected as shown. It's a PNP transistor, characterized
by its three terminals: collector, base, and emitter. The direction of the
current is from the emitter to the collector, which is a characteristic of
PNP transistors. In PNP transistors, the majority of charge carriers
responsible for the current are holes.
Let's label the terminals: 'p' for the emitter, 'n' for the base, and 'p' for the
collector. We have an external voltage source, 'Vi,' providing an AC
alternating current signal applied across these two terminals. 'Vo'
represents the output voltage, which is also an AC signal taken across
these two terminals.
To provide biasing to this BJT (Bipolar Junction Transistor) circuit and
set its operating point, we have two voltage sources, 'VEE' and 'VCC,'
connected via resistors, as shown. When we refer to DC biasing, we
mean setting the operating point or quiescent point (Q-point) around
which the externally applied AC signal will be processed, effectively
amplifying it.
Let's examine the biasing conditions:
1. The positive terminal of 'VEE' connects to the emitter through
the resistor 'RE,' and the negative terminal goes to the base.
This forward-biases the emitter-base junction, creating a
voltage 'VBE' across this junction.
2. The negative terminal of 'VCC' connects to the collector (P-type
material), while the positive terminal goes to the base (N-type
material) through the resistor 'RC.' This arrangement reverse-
biases the base-collector junction, resulting in a voltage 'VCB'
across this junction.
With the emitter-base junction forward-biased and the base-collector
junction reverse-biased, the transistor is biased in the active region,
allowing it to function as an amplifier. This is how the circuit operates.
Transistor in common collector
mode | Emitter follower | voltage
follower
Figure 6‑8 BJT in CC mode [55]
In this circuit example (Figure 6‑8), we observe an NPN transistor. The
direction of conventional current indicates that it flows out of the
emitter. This characteristic identifies it as an NPN transistor, where the
current flows from the collector to the emitter.
The three terminals of the transistor are the collector, base, and emitter.
The N-region, P-region, and N-region arrangement confirms that this is
an NPN transistor.
The input signal is applied to the base terminal, and the output is taken
from the emitter. This configuration is known as the common collector
(CC) configuration, where the collector is connected to the positive
supply voltage.
The purpose of this common collector configuration is to bias the
transistor at a DC operating point using external resistors (not shown).
This biasing ensures that the base-to-emitter junction is forward-biased,
while the base-to-collector junction is reverse-biased. Such biasing
places the transistor in the active region.
When an external AC input signal is applied to the base terminal, it
results in base current flowing into the base terminal. The emitter current
(IE) and collector current (IC) flow accordingly.
Neglecting the base current contribution, we can state that IE equals IC.
Considering Kirchhoff's voltage law (KVL) around the loop, we have Vin
- VBE - IE × RE = 0.
By substituting IE × RE with Vout, we find that Vout = Vin - VBE. This
equation demonstrates that the output voltage in this circuit will always
be equal to the input voltage minus the base-emitter junction voltage
drop, typically around 0.6 to 0.7 volts.
This circuit is known as a voltage follower circuit, which is a common
application of a BJT biased in common collector mode. It is also referred
to as an emitter follower because the voltage at the emitter closely tracks
the voltage at the base.
How will transistor maintain the
constant bias in this circuit?
Let's revisit the circuit in Figure 6‑8 in our previous module. As
observed, the output (Vo) closely tracks the input (Vin), where the voltage
drop VBE is consistently maintained across the base-emitter junction.
To understand this circuit, let's consider a scenario where there is a
variation in the input voltage (Vin), either increasing or decreasing. For
instance, if Vin increases, the base current (IB) of the transistor also
increases due to the transistor's action. The collector current (IC) is the
product of beta (β) multiplied by IB, where β represents the current gain
of the transistor. Consequently, with increased Vin, IB and IC increase, and
the collector current approximately equals the emitter current (IE) that
flows through the resistor RE.
As a result, the voltage drop across RE, which is equivalent to Vout,
increases. Therefore, when Vin rises, Vout endeavors to catch up, and this
relationship is proportional.
Conversely, when Vin decreases, IB and IC decrease, causing a reduction
in IE, which in turn leads to a decrease in the voltage drop across RE,
resulting in a lower Vout. Hence, for both increases and decreases in Vin,
Vout adapts proportionately.
This phenomenon ensures that the voltage drop VBE remains constant.
The voltage at the base is at the positive terminal, while the emitter
voltage is at the negative terminal. This constant VBE drop is maintained,
allowing the transistor to maintain equilibrium in the circuit.
It's worth noting that variations in the base-emitter voltage (VBE) may
occur due to changes in the DC biasing. In such cases, the transistor
responds by adjusting the voltage at the emitter accordingly, ensuring
that the VBE drop is consistently maintained. This circuit configuration is
known as the common collector configuration, where the output closely
follows the input.
Fixed bias or base bias - 1
Figure 6‑9 Fixed bias circuit (Base bias) [56]
In this circuit (Figure 6‑9), we're examining one of the fundamental
biasing techniques for Bipolar Junction Transistors (BJTs). This
configuration is known as the fixed bias or base bias circuit, often
referred to as base bias because it uses fixed resistor values (RB and RC)
to establish the DC operating point or Q point of the circuit. Let's delve
into how it operates.
In this setup, we have an NPN transistor with three terminals: collector,
base, and emitter. The voltage across the base and emitter is termed VBE,
while the voltage across the collector and emitter is denoted as VCE.
For the input side, KVL yields: VCC - IBRB - VBE = 0
For the output side, KVL yields: VCC - ICRC - VCE = 0
From these equations, we can compute IB, which is related to IC through
the transistor's current gain (β) with the equation IC = β * IB. This base
current, IB, enables us to determine IC and VCE.
It's worth noting that in this configuration, the base-emitter junction is
forward-biased, while the collector-base junction is reverse-biased,
placing the transistor in the active region of operation.
Now, let's discuss the drawbacks of this circuit. Firstly, as the
temperature increases, causing the collector current (IC) to rise, the base
current (IB) also increases due to the transistor equation. This shift in IC
and IB alters the original operating point, impacting circuit stability.
Secondly, variations in the transistor's current gain (β) can result in
different operating points, further affecting circuit stability. To mitigate
these stability issues related to temperature and β variations,
modifications like adding an emitter resistor (RE) in the emitter path are
often employed.
Fixed bias or base bias - 2
Let's take a closer look at this Bipolar Junction Transistor (BJT) circuit.
This circuit is known as fixed bias with emitter resistor or fixed bias
with emitter degeneration resistor. The emitter resistor is positioned here
to enhance circuit stability and reduce harmonic distortion.
As you can observe, it features an NPN transistor with collector, base,
and emitter pins. The base-emitter voltage is denoted as VBE, and the
collector-emitter voltage is VCE. The term "fixed bias" originates from
the fact that we apply a supply voltage, VCC, and connect a resistor to the
base terminal, which establishes biasing using the supply voltage itself.
In this configuration, two Kirchhoff's voltage law equations can be
applied:
1. VCC - IBRB - VBE - IERE = 0 Here, IB represents the current
through the base resistor, VBE is the voltage across the base-
emitter junction, and IERE signifies the emitter current.
2. VCC - ICRC - VCE - IRB - IERE = 0 This equation accounts for the
current through the collector resistor (RC), the collector-emitter
voltage (VCE), the current through the base resistor (RB), and
the emitter resistor (RE).
Figure 6‑10 Fixed bias with emitter resistor [56]
This biasing arrangement is termed "fixed bias" because biasing is
achieved through a fixed resistor, RB, through which the base current
flows. The fundamental operation of this NPN BJT transistor remains
consistent in this configuration.
collector base bias or collector
feedback biasing
Figure 6‑11 Collector-to-base bias [56]
This is another circuit used for biasing the BJT transistor, known as the
collector-to-base bias circuit (Figure 6‑11). In this configuration, a
supply voltage VCC is applied, and we have resistors RC and RB
connected to the circuit. This arrangement earns it the name "collector-
to-base register biasing" for the BJT.
As with most circuits, this one involves an NPN transistor with collector,
emitter, and base pins. The base-emitter junction is forward-biased,
while the base-collector junction is reverse-biased, setting the transistor's
operating point in the active region. Resistors RB and RC are employed to
establish the circuit's operating point.
When the supply voltage is activated, current flows into the circuit,
denoted as "I." This current splits into two branches, one being IC and the
other IB. So, essentially, the total current I is the sum of IC and IB.
Applying Kirchhoff's Voltage Law (KVL) to the circuit, we have VCC -
voltage drop across RC - voltage drop across RB - VBE = 0. Utilizing this
KVL equation allows us to calculate current and voltage values within
this circuit loop.
RB and RC play a crucial role in setting the circuit's operating point.
There are two primary factors that can change the operating point, and
we will examine how this circuit helps stabilize it.
Firstly, changes in temperature can alter the operating point. For a fixed
VBE, if the circuit's temperature increases, the collector current starts to
rise. Since collector current is related to base current through the
equation IC = β * IB, an increase in IC leads to an increase in IB. This, in
turn, changes the operating point due to temperature rise.
The circuit responds to the temperature rise by causing the voltage drop
across RC to increase because of the increasing IC. Consequently, the
voltage drop across RB decreases. According to KVL, VRB (voltage
across RB) decreases, causing IB to decrease. This phenomenon, termed
negative feedback, counteracts the initial increase in IB due to rising
temperature.
Secondly, transistor beta (β) variations can also affect the operating
point. Replacing the transistor in the circuit with one having a different β
will result in a different IC and IB. However, by connecting resistor RB
from the collector to the base, the circuit mitigates the impact of β
variations on the DC operating point, providing stability in the face of
changes in temperature or transistor beta.
In summary, this collector-to-base bias circuit is designed to stabilize the
operating point against variations in temperature and transistor beta,
ensuring consistent performance under different conditions.
Voltage divider bias circuit for
transistor | BJT biasing
Figure 6‑12 BJT voltage divider bias [56]
Here is one of the most popular biasing techniques for bipolar junction
transistors, known as the voltage divider bias circuit (Figure 6‑12). In
this configuration, we have an NPN transistor with three pins: collector,
base, and emitter. The voltage across the base and emitter is referred to
as VBE, and the voltage across the collector and emitter is denoted as VCE.
To set the DC operating point for this transistor, we use two resistors, R1
and R2, along with a supply voltage, VCC. The voltage division created
by R1 and R2 applies a voltage across the base-emitter junction, which
can be calculated as follows:
VR2 = VCC * (R2 / (R1 + R2))
This is a simple application of the potential divider rule. The resulting
voltage, VR2, is applied across the base-emitter junction. Next, there is an
emitter current (IE) that flows through the resistor RE. To analyze this
circuit, we can apply Kirchhoff's voltage law (KVL) to the loop
involving VR2, VBE, and IE.
Similarly, for the outer circuit of the BJT, we can apply KVL to the loop
involving VCC, ICRC (collector current through RC), and VCE. These two
loops help us determine the voltages and currents in the circuit.
The base current (IB) can be calculated as the difference between I1 and
I2:
IB = I1 - I2
This voltage division circuit allows us to establish a DC operating point
for the transistor.
Please note that the specific values of currents and voltages would
depend on the component values and transistor characteristics used in
the circuit.
Chapter 7 Principles of electronic Measurements
Measuring voltage, current, and resistance
How to measure resistance using
digital multi-meter (DMM)?
Figure 7‑1 Resistance measurement using DMM
Measuring resistance using a Digital Multi meter (DMM) is a
fundamental and straightforward process commonly used in electronics
and electrical troubleshooting. Here are the steps to measure resistance
using a DMM:
Steps to Measure Resistance with a DMM:
1. Turn On the DMM:
Power on the DMM by rotating the mode dial or
pressing the power button, depending on the
model.
2. Select Resistance (Ω) Mode:
Set the DMM to the resistance (Ω) measurement
mode. This mode is usually indicated by the Greek
letter omega symbol (Ω) on the DMM's mode dial.
3. Choose the Range:
If your DMM offers multiple resistance
measurement ranges, select a range that is equal to
or greater than the expected resistance of the
component you are testing. It's best to start with
the highest range and work your way down for
accuracy.
4. Connect the Test Leads:
Connect the black test lead (negative or common)
to the common (COM) terminal on the DMM.
Connect the red test lead (positive) to the terminal
labeled for resistance measurements (Ω or
resistance).
5. Measure the Resistance:
Place the test leads across the component or
resistor you want to measure. For through-hole
resistors, touch the leads to each resistor terminal.
For SMD resistors, use tweezer probes or fine-
point probes to make contact with the resistor's
terminations.
6. Read the Measurement:
The DMM will display the resistance value on its
screen. If you're using a digital multimeter, it will
show the value in ohms (Ω). If you're using an
analog multimeter, read the resistance value from
the scale and consider any multiplier markings
(e.g., kΩ for kilohms).
7. Record the Value:
Note down the measured resistance value for
reference.
8. Disconnect the Test Leads:
Remove the test leads from the component you've
measured.
9. Turn Off the DMM:
To conserve battery life, turn off the DMM when
you've completed your measurements.
Chapter 8 Additional topics
Integrated circuits (ICs)
Figure 8‑1 A microscope image of an integrated circuit (IC) die. The
pinouts are the black circles surrounding the IC [57].
Figure 8‑1 shows us a micrograph of an IC, which stands for an
integrated circuit. An IC stands for "integrated circuit." This micrograph
depicts a piece of silicon. Integrated circuits are typically made from
semiconductor material, such as silicon, which is clearly visible here.
This is also commonly referred to as a "chip," "die," or "IC die." When
examining the details and microfeatures of the electronic components
integrated onto this die, a microscope is required.
Usually, the dimensions of these microchips are in the range of just one
or two square millimeters. For example, the height and width of such a
chip can be as small as 1 mm x 1 mm.
The central area of this chip is where various electronic components,
devices, and circuits have been integrated. These chips can serve various
purposes, such as microprocessors, microcontrollers, memory, analog
circuits, radio frequency circuits, and other digital logic components, all
compactly integrated within this small area.
The black circles located around the perimeter of the IC represent the
pinout. They indicate the connection points for the circuits within. In the
modern era of microelectronics, electronic components have been
miniaturized and integrated onto pieces of silicon, leading to a
significant reduction in the cost of electronics. Microprocessors or
microcontroller ICs, for instance, are now available at very affordable
prices, making them accessible to a broader audience.
The development, manufacturing, and testing of integrated circuits
require a range of skills and expertise. Designing, producing, and
ensuring the quality of these ICs are essential steps in the electronics
industry.
Analog and Digital | electronic | basics and advance concepts
Figure 8‑2 Analogue and digital world [58]
In Figure 8‑2, we have a detailed illustration explaining a fundamental
concept in electronics – the transmission and reception of electronic
signals, particularly in the domain of acoustics or sound. This serves as
an excellent opportunity to grasp the difference between analog and
digital signals and understand how these signals are processed, including
the instruments and devices used to convert one form of energy into
another. Central to this process are transducers or sensors.
To the far left in the image, you'll notice someone speaking. This spoken
audio is represented by sound waves, which can be considered acoustic
signals. These signals possess specific amplitudes and frequencies,
typically ranging from 20 Hertz to 20 kilohertz. To process these sound
waves or acoustic signals, we employ an electronic device known as a
microphone, which acts as a transducer. It captures the sound vibrations
and converts them into proportional electrical voltages. This
transformation is depicted in the accompanying graph, demonstrating an
analog signal with electrical voltage that continuously varies with
respect to time. It exhibits both positive and negative amplitudes,
creating an instantaneous signal concerning time.
Now, turning our attention to the rectangular box, we find a computer.
Within the computer, essential components include an Analog-to-Digital
Converter (ADC), a Digital-to-Analog Converter (DAC), and digital
circuits. The ADC primarily converts the analog voltage, as seen in the
graph, into digital codes, effectively transitioning from the analog
domain to the digital domain. These digital codes are then subjected to
further processing using Digital Signal Processing (DSP) circuitry,
which may involve various operations like filtering, amplification, and
modulation.
The digital codes generated during this processing phase are then sent to
the Digital-to-Analog Converter (DAC), which performs the reverse
transformation, converting digital codes back into proportional electrical
voltages. In this manner, the signal transitions from the digital domain
back to the analog domain, and we have an analog signal, as portrayed in
the second graph.
Finally, this analog signal is fed to another transducer, a loudspeaker,
which converts the incoming electrical signal into sound waves
proportional to the original acoustic signal. These sound waves are then
presented to the human ear for audible reception.
In this manner, the complete communication system operates – from the
source (such as a speaker) to the receiver (the listener). Electronics plays
a pivotal role in enabling this transmission of sound and information. We
hope you found this explanatory topic informative and appreciate the
role of technology in transducing analog and digital signals.
What is noise in Electronics? What causes it?
In this module, we take a closer look at Figure 8‑3 to understand the
concept of noise in typical electronic systems and electronic signals. In
electronics, there are two types of signals: analog and digital. Let's start
by understanding what an analog signal is. On the y-axis, the signal can
represent current, voltage, or a combination of both, and it continuously
varies with respect to time. On the other hand, a digital signal switches
between two levels, often referred to as high and low levels or positive
and negative levels. These two logic levels define digital signals.
Figure 8‑3 Electronic Noise [59]
Now, let's get back to our main topic, which is noise. Whether you have
an analog or digital signal, noise is always present in electronic systems.
The components and devices used to process these signals, whether
analog or digital, produce fluctuations due to the atomic structures of
these components. These atomic fluctuations are introduced into the
system.
When noise is added to the original signals, the signal may look like this.
In the case of an analog signal, the original signal gets mixed with noise
and may appear corrupted. Similarly, with a digital signal, the original
signal may look like this when noise is present.
In practical terms, think of it as an example: when I'm explaining this
concept to you through my voice, my original signal, which is my voice,
is reaching you. However, in the background, there's additional audio,
which we can refer to as noise, mixing with my voice. This illustrates
how original signals can get corrupted when noise is introduced into the
system.
Hope you liked the contents of this book, if you did so, do share with others
and visit our website for business enquiries and communication.
Visit our YouTube channel and subscribe it for updates and valuable
contents.
References
[1] “No Title.”
https://en.wikipedia.org/wiki/Electronics#/media/File:Componentes.J
PG
[2] “No Title.”
https://commons.wikimedia.org/wiki/File:Diod_LED_symbol.png
[3] “No Title”, [Online]. Available:
https://commons.wikimedia.org/wiki/File:Carbon_Composition_Resi
stor_1K3_cracked.png
[4] “No Title”, [Online]. Available:
https://commons.wikimedia.org/wiki/File:10k_Ohm_Potentiometer_-
_POT-103-BIG_%284149943231%29.jpg
[5] “No Title”, [Online]. Available:
https://commons.wikimedia.org/wiki/File:10k_Ohm_Breadboard_Co
mpatible_Potentiometer.jpg
[6] “No Title”, [Online]. Available:
https://commons.wikimedia.org/wiki/File:100_ohm_SMD_1206_resi
stor.jpg
[7] “No Title”, [Online]. Available:
https://commons.wikimedia.org/wiki/File:22_ohm_SMD_0603_resist
or.jpg
[8] “No Title”, [Online]. Available:
https://www.digikey.in/en/resources/conversion-
calculators/conversion-calculator-smd-resistor-code
[9] “webpage.” https://www.utmel.com/tools/smd-resistor-code-
calculator?id=33
[10] “No Title.”
https://commons.wikimedia.org/wiki/File:VFPt_capacitor_mod.svg
[11] “No Title.”
https://commons.wikimedia.org/wiki/File:Electronic_component_cap
acitors.jpg
[12] “No Title.” https://creativecommons.org/licenses/by-sa/4.0/
[13] “No Title.” https://creativecommons.org/licenses/by-sa/2.0/
[14] “No Title.” https://creativecommons.org/licenses/by/2.0/
[15] “No Title.” https://creativecommons.org/licenses/by-sa/3.0/
[16] “No Title.”
https://commons.wikimedia.org/wiki/File:Inductor_Symbols.jpg
[17] “No Title.” https://en.wikipedia.org/wiki/Alternating_current
[18] “No Title.”
https://commons.wikimedia.org/wiki/File:Conventional_Current.png
[19] “No Title”, [Online]. Available:
https://en.wikipedia.org/wiki/File:Non-
ideal_voltage_and_current_sources.svg
[20] “No Title”, [Online]. Available:
https://commons.wikimedia.org/wiki/File:Block_resistance.svg
[21] “No Title”, [Online]. Available:
https://commons.wikimedia.org/wiki/File:Ohm%27s_law_triangle_%
28VIR%29.jpg
[22] “No Title.” https://en.wikipedia.org/wiki/Kirchhoff’s_circuit_laws
[23] “No Title.” https://et.wikipedia.org/wiki/Kirchhoffi_seadused
[24] “No Title.”
https://commons.wikimedia.org/wiki/File:Kirchhoff%27s_current_la
w.jpg
[25] “No Title.” https://en.wikipedia.org/wiki/Series_and_parallel_circuits
[26] “No Title.” https://bn.wikipedia.org/wiki/ #/media/
:Series_resistors.jpg
[27] “No Title.”
https://su.wikipedia.org/wiki/Résistor#/media/Gambar:Resistorscomb
o.png
[28] “No Title.”
https://en.wikipedia.org/wiki/Reflections_of_signals_on_conducting_
lines
[29] “No Title.”
https://commons.wikimedia.org/wiki/File:Voltage_divider-loaded.svg
[30] “No Title.” https://en.wikipedia.org/wiki/Current_divider
[31] “No Title.”
https://commons.m.wikimedia.org/wiki/File:Energy_bands_of_a_sem
iconductor_type_N.svg
[32] “No Title.” https://commons.wikimedia.org/wiki/File: .png
[33] “No Title.” https://commons.wikimedia.org/wiki/File:SY320-
10_HFO.jpg
[34] “No Title.”
https://commons.wikimedia.org/wiki/File:Germanium_Diode_1N60P
.jpg
[35] “No Title.” https://commons.m.wikimedia.org/wiki/File:V-
a_characteristic_diodes_si_ge.svg
[36] “No Title.”
[37] “No Title.”
https://commons.m.wikimedia.org/wiki/File:Transistors.agr.jpg
[38] “No Title.”
https://commons.wikimedia.org/wiki/File:NPN_BJT_Basic_Operatio
n_%28Active%29_fr.svg
[39] “No Title.”
https://commons.wikimedia.org/wiki/File:NPN_AND_PNP_BJT_SY
MBOLS.png
[40] “No Title.”
https://upload.wikimedia.org/wikipedia/commons/9/91/Transistor_Si
mple_Circuit_Diagram_with_NPN_Labels.svg
[41] “No Title.” https://en.wikipedia.org/wiki/Field-effect_transistor
[42] “No Title.” https://et.wikipedia.org/wiki/Pn-siirdega_väljatransistor
[43] “No Title.”
https://en.wikipedia.org/wiki/List_of_MOSFET_applications
[44] “No Title.”
https://commons.wikimedia.org/wiki/File:FET_Symbols.svg
[45] “No Title.”
https://en.wikipedia.org/wiki/File:MOSFET_transistors.jpg
[46] “No Title.”
https://en.wikipedia.org/wiki/File:Scheme_of_metal_oxide_semicond
uctor_field-effect_transistor.svg
[47] “No Title.”
https://en.wikipedia.org/wiki/CMOS#/media/File:Cmos_impurity_pr
ofile.PNG
[48] “No Title.”
https://commons.wikimedia.org/wiki/File:SE_Half_Wave_Rectifier.s
vg
[49] “No Title.”
https://commons.wikimedia.org/wiki/File:SE_Full_wave_rectifier_-
_Center_tapped.svg
[50] “No Title.” https://commons.wikimedia.org/wiki/File:RC_filter.svg
[51] “No Title.” https://en.wikipedia.org/wiki/Common_emitter
[52] “No Title.”
[53] “No Title.”
https://commons.m.wikimedia.org/wiki/File:NPN_BJT_common_bas
e_biasing-en.svg
[54] “No Title.”
https://commons.m.wikimedia.org/wiki/File:PNP_BJT_common_Bas
e_bias.svg
[55] “No Title.” https://en.m.wikipedia.org/wiki/Common_collector
[56] “No Title.” https://en.wikipedia.org/wiki/Bipolar_transistor_biasing
[57] “No Title.” https://en.wikipedia.org/wiki/Integrated_circuit
[58] “No Title.” https://en.wikibooks.org/wiki/A-
level_Computing/AQA/Paper_2/Fundamentals_of_data_representati
on/Analogue_and_digital
[59] “No Title.” https://commons.m.wikimedia.org/wiki/File:Electronics-
noise-1-638.jpg
[60] “No Title.” https://en.wikipedia.org/wiki/Electronics