Ee 211 Short
Ee 211 Short
Ee 211 Short
Hayrettin Köymen
Abdullah Atalar
Bilkent University
1
ISBN: 978-605-9788-47-2
Contents
Preface vi
3 AUDIO CIRCUITS 77
3.1 Complex numbers . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.2 Phasors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
3.2.1 Derivative operator . . . . . . . . . . . . . . . . . . . . . . 81
3.2.2 Integration operator . . . . . . . . . . . . . . . . . . . . . 81
3.2.3 Resistor with sinusoidal excitation . . . . . . . . . . . . . 81
3.2.4 Capacitor with sinusoidal excitation . . . . . . . . . . . . 81
3.2.5 Inductor with sinusoidal excitation . . . . . . . . . . . . . 82
3.3 Linear circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
3.3.1 Steady-state solution of linear RLC Circuits with sinu-
soidal excitation . . . . . . . . . . . . . . . . . . . . . . . 83
3.3.2 Power relation for phasors . . . . . . . . . . . . . . . . . . 85
3.3.3 Impedance . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.3.4 Admittance . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.4 Transfer function . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
3.4.1 Transfer function of first-order circuits . . . . . . . . . . . 89
3.4.2 Transfer function of higher-order circuits . . . . . . . . . . 91
3.5 Thévenin Equivalent Circuit . . . . . . . . . . . . . . . . . . . . . 92
3.5.1 Thévenin equivalent circuit for RLC networks with sinu-
soidal excitation . . . . . . . . . . . . . . . . . . . . . . . 94
3.6 Norton Equivalent Circuit . . . . . . . . . . . . . . . . . . . . . . 95
3.6.1 Norton equivalent circuit for RLC networks with sinu-
soidal excitation . . . . . . . . . . . . . . . . . . . . . . . 97
3.6.2 Using Thévenin and Norton equivalent circuits . . . . . . 98
3.7 Superposition principle . . . . . . . . . . . . . . . . . . . . . . . . 100
3.8 Amplifier types . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
3.9 Operational amplifiers . . . . . . . . . . . . . . . . . . . . . . . . 104
3.9.1 Inverting amplifier . . . . . . . . . . . . . . . . . . . . . . 105
3.9.2 Non-inverting amplifier . . . . . . . . . . . . . . . . . . . 106
3.9.3 Summing amplifier . . . . . . . . . . . . . . . . . . . . . . 106
3.9.4 Difference amplifier . . . . . . . . . . . . . . . . . . . . . . 107
CONTENTS iii
4 DIODES and
BIPOLAR JUNCTION TRANSISTORS 134
4.1 Diodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
4.1.1 Schottky diode . . . . . . . . . . . . . . . . . . . . . . . . 136
4.1.2 Solutions of circuits containing diodes . . . . . . . . . . . 136
4.1.3 Diodes as rectifiers . . . . . . . . . . . . . . . . . . . . . . 140
4.1.4 Half-wave rectifier . . . . . . . . . . . . . . . . . . . . . . 140
4.1.5 Full-wave rectifier . . . . . . . . . . . . . . . . . . . . . . 145
4.1.6 Bridge rectifier . . . . . . . . . . . . . . . . . . . . . . . . 147
4.1.7 Zener diodes as voltage sources . . . . . . . . . . . . . . . 148
4.1.8 LED . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
4.2 Bipolar Junction Transistor (BJT) . . . . . . . . . . . . . . . . . 152
4.2.1 States of a BJT . . . . . . . . . . . . . . . . . . . . . . . . 153
4.3 DC Analysis of BJT circuits . . . . . . . . . . . . . . . . . . . . . 154
4.4 Biasing of BJTs . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
4.4.1 Simple Base Bias . . . . . . . . . . . . . . . . . . . . . . . 157
4.4.2 Base Bias with Emitter Resistor . . . . . . . . . . . . . . 158
4.4.3 Conventional Bias Circuit . . . . . . . . . . . . . . . . . . 159
4.4.4 Bias Circuit with Collector Feedback . . . . . . . . . . . . 160
4.5 Small-signal BJT Model . . . . . . . . . . . . . . . . . . . . . . . 162
4.6 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
4.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
6 FILTERS 220
6.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
6.1.1 Transducer power gain of a filter . . . . . . . . . . . . . . 221
6.1.2 First- and second-order low-pass-filters . . . . . . . . . . . 221
6.2 Polynomial filters . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
6.2.1 Butterworth filters . . . . . . . . . . . . . . . . . . . . . . 224
6.2.2 Chebyshev filters . . . . . . . . . . . . . . . . . . . . . . . 230
6.2.3 Practical aspects of LC filter design . . . . . . . . . . . . 232
6.3 Impedance matching . . . . . . . . . . . . . . . . . . . . . . . . . 233
6.3.1 Matching by transformers, narrow-band . . . . . . . . . . 233
6.3.2 Matching by transformers, wide-band . . . . . . . . . . . 235
6.3.3 Matching by resonant circuits . . . . . . . . . . . . . . . . 236
6.3.4 Impedance inverters . . . . . . . . . . . . . . . . . . . . . 237
6.3.5 Band-pass-filter design using inverters . . . . . . . . . . . 239
6.4 Crystal filters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241
6.4.1 Band-pass-filter using quartz crystals . . . . . . . . . . . . 242
6.5 SAW filters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
6.6 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
6.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
7 DIODES IN
TELECOMMUNICATIONS 255
7.1 Envelope detector . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
7.1.1 Real diodes in envelope detectors . . . . . . . . . . . . . . 257
7.2 Automatic gain control . . . . . . . . . . . . . . . . . . . . . . . . 259
7.2.1 PIN diode . . . . . . . . . . . . . . . . . . . . . . . . . . . 259
7.2.2 Automatic gain control using a PIN diode . . . . . . . . . 260
7.3 Signal presence indicator . . . . . . . . . . . . . . . . . . . . . . . 261
7.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
Bibliography 318
Index 320
Preface
grated circuits, as active devices, are discussed. Filters, power supplies, audio
amplifiers, speakers, microphones, radio amplifiers, oscillators, mixers, inter-
modulation, and antennas are progressively introduced towards the construc-
tion of the transceiver. Minimum mathematical background and definitions
(such as the solution of first-order differential equations and phasors) are intro-
duced only when necessary. All terminology and jargon are introduced. A PCB
of the transceiver is provided with the course kit and components.
Another aspect of the course is that every student must possess a soldering
iron, de-soldering pump, a multimeter, a scientific calculator and a set of hand
tools. Possessing such electronics-specific tools improves the ties and commit-
ment to the discipline and the enthusiasm to learn electronics. Compared to
EE education thirty years ago, the students suffer an identity problem today.
Thirty years ago, every engineering student had to have a set of drafting tools
and a slide rule right at the beginning of freshman. Computers replaced these
today. Computers, however, are anonymous. Students of almost every discipline
use computers. Computers are not specific to electronics students.
These considerations made this book different from more conventional elec-
tronics textbooks. The book tells the story of making a transceiver and intro-
duces various concepts and other information only when necessary. In other
words, the related topics in a subject are not, generally, collected together in
the same section in this book. They are given at the relevant stages of the
transceiver construction.
Acknowledgements:
Many people contributed both to design the course and to the course mate-
rial. The author wishes to acknowledge the tremendous effort that Müjdat Bal-
antekin had put into this work. The laboratory material for this course would
have never been realized without the support provided by Ergün Hırlakoğlu,
İsmail Kır and Ersin Başer. One hundred students of the class of 2001 provided
invaluable feedback. The author acknowledges their contributions, effort and
positive energy. Prof. A. Altıntaş, Prof. A. Atalar, Prof. B. Özgüler, Dr. T.
Reyhan, Dr. S. Topçu, Dr. E. Tın, Prof. C. Yalabık, of Bilkent University and
E. Ceyhan of ERE Corp. made many suggestions and critically reviewed the
text. Finally, the author is indebted to N. Özönder of Telmek Corp. and B.
Arıkan of Arıkan Elektronik, for an immaculate TRC-10 instrument tray and
PCB.
Hayrettin Köymen
August 25, 2002
Ankara
PREFACE viii
Abdullah Atalar
July 2023
Ankara
Chapter 1
SIGNALS AND
COMMUNICATIONS
1.4 Wavelength
Electromagnetic waves travel at the speed of light, c = 3.0 × 108 m/sec in the
air (or vacuum). This speed is the same as the speed of light in air, since light
1.4. WAVELENGTH 3
Table 1.1: Frequencies and wavelengths of light waves. Refer to page 310 for
the definition of unit prefixes.
c
λ= (1.3)
f
where λ is the wavelength in meters. Table 1.1 shows the frequency and wave-
length of light waves.
Radio waves used in electronic communication are also electromagnetic waves.
Table 1.2 lists some commonly used frequencies and the corresponding wave-
lengths.
Fig. 1.2 is a diagram showing the frequency spectrum with labels given to
different bands and the corresponding wavelengths.
The formula of Eq. 1.3 also applies to sound waves, which travel in a medium.
Sound waves propagate at a speed of 340 m/s in air and 1500 m/s in water.
Therefore, a sound wave at a frequency of 1.0 kHz has a wavelength of 34 cm
in air and 1.5 m in water.
1.5. OSCILLATORS 4
µm
nm
km
λ
m
m
10 km
10 cm
1 mm
1 µm
10 m
1 km
1 cm
10 µ
100
100
1m
100
100
VLF LF MF HF VHF UHF SHF EHF FIR MIR NIR NUV
3 kH
30 kH
300
3 MH
30 M
300
3 GH
30 G
300
3 TH
30 T
300
3 PH
f kHz
MHz
GHz
THz
Hz
z
Hz
z
Hz
z
z
z
z
TRC-11
1880
1700 z
108
174
1710
26.1 z
88 M
880 MHz
960 Hz
535
5.9 M
27 M Hz
54 M
216
2.4 G z
2.5 GHz
470
806
MHz
MHz
M
MHz
kH
MHz
MHz
M
Hz
Hz
Hz
MH
kHz
MHz
Hz
H
Figure 1.2: The frequency spectrum and its bands: VLF (very low frequency),
LF (low frequency), MF (medium frequency), HF (high frequency), VHF (very
high frequency), UHF (ultra high frequency), SHF (super high frequency), EHF
(extremely high frequency), FIR (far infrared), MIR (mid infrared), NIR (near
infrared). Some bands reserved broadcast radio transmission (AM, SW and
FM), TV transmission, cellular phone (GSM) and local-area wireless (Wi-Fi)
are also shown.
1.5 Oscillators
Electronic circuits that generate voltages of sinusoidal waveform are called si-
nusoidal oscillators. There are also oscillators generating periodic signals of
other waveforms, among which square wave generators are the most popular.
Square wave oscillators are predominantly used in digital circuits to produce
time references, synchronization, etc. For example, an electronic wristwatch
has an oscillator at a frequency of 32768 Hz, which is easily divided to 215 using
a 15-stage divide-by-two circuit to generate 1.0000 pulses per second* . Function
generators are capable of producing a number of periodic signals like sinusoidal,
square, triangular and sawtooth waveforms, with frequencies and amplitudes
adjustable by the front panel buttons (see for example Fig. 1.3).
We use sinusoidal oscillators in communication circuits for various reasons.
In most cases, the oscillators determine the frequency of operation.
* 1.0000 expresses the precision of the quantity of one pulse per second. Refer to Appendix B
for significant figure notation.
1.5. OSCILLATORS 5
2
amplitude (V)
0 T/2 T
time
2.5
a0 +Fundamental
2 Up to 3rd Harmonic
Up to 7th harmonic
Up to 59th harmonic
1.5
0.5
-0.5
0 T/5 2T/5 3T/5 4T/5 T 6T/5
Figure 1.5: Constructing a square wave from harmonics, (a) only a0 + funda-
mental, (b) all terms up to 3rd harmonic, (c) all terms up to 7th harmonic, (d)
all terms up to 59th harmonic.
reasonably well delineated. It looks more like a square wave as the number of
added harmonics increase.
A common graphical representation of a signal with many sinusoidal com-
ponents is to plot the line graph of the amplitude of each component versus
frequency (either f or ω). This is called the spectrum of the square wave or its
frequency domain representation. The spectrum of this square wave is given
in Fig. 1.6, which clearly illustrates the frequency components of the square
wave. The figure shows that the square wave, being a periodic signal, has en-
ergy only at discrete frequencies, more specifically only at the odd harmonics of
the fundamental.
♦ TRC-11 has two oscillators, one at the frequency of 12.00 MHz, and the
other at 15.00 MHz.
1.6. MODULATION 7
4/π
|bn|
4/3π
4/5π
4/7π
0 ω 3ω 5ω 7ω
Frequency
1.6 Modulation
We frequently use electromagnetic waves to transmit information from one place
to another. The information, for example, voice or music, must first be converted
into an electrical voltage, vm (t). We, then, convert the electrical signal to an
electromagnetic wave to transmit it over some distance to the receiver. The
conversion of vm (t) to an electromagnetic signal occurs by using an antenna.
We will see in Chapter 8 that the size of the antenna should be comparable to
the wavelength of the signal for efficient conversion.
The wavelength of an electrical voice or music signal is measured in hundreds
of kilometers. Using an antenna of that size is obviously not practical. To
make the antenna size small, we need to use a much higher frequency sinusoid
(called the carrier) with a much smaller wavelength to carry the information. In
order to transmit voice or music, we need to make one parameter of this carrier
sinusoid dependent on the information. Merging the information-carrying signal
on a high frequency carrier sinusoid is called modulation.
There are three parameters that we can modify in a sinusoid: amplitude,
frequency and phase. In the amplitude modulation (AM) method, the amplitude
of a sinusoid is made dependent on vm (t). Let us assume that vm (t) is a simple
signal, Vm cos(ωm t). In order to modulate the amplitude of a carrier signal,
Vc cos(ωc t), we construct the signal,
Vm
v(t) = Vc cos(ωc t) + vm (t) cos(ωc t) = Vc 1 + cos(ωm t) cos(ωc t) (1.5)
Vc
vm (t) is called the modulating signal. In AM, the maximum peak variation
of |vm (t)| must always be less than Vc , otherwise some parts of vm (t) get lost.
Vc [1+(Vm /Vc ) cos(ωm t)] part in AM signal is called the envelope. An AM signal
is depicted in Fig. 1.7(b).
The depth of modulation is determined by the maximum value of the nor-
malized modulation signal |vm (t)/Vc |. The modulation index, m, is defined as
Vm
m= (1.6)
Vc max
(c)
Figure 1.7: (a) Modulating signal, vm , (b) AM modulated signal, (c) FM mod-
ulated signal.
such that β = kf /ωm and ω(t) = dθ(t)/dt = ωc + kf vm (t). Here, we change the
instantaneous frequency of the carrier signal around the carrier frequency, ωc ,
according to the variation of modulating (information) signal, while the envelope
of the signal stays constant. An FM modulated signal is shown in Fig. 1.7(c).
Long wave and middle wave radio broadcasting are done by AM, and radio
broadcasting in the 88–108 MHz band is done by FM. Analog terrestrial televi-
sion broadcasting employs a version of AM (called vestigial side-band AM) for
image and FM for sound.
1.7 Amplifiers
The most frequently done operation on signals is amplification. The signal
received at an antenna is often very weak, may be at power levels of a few tens
of fW (1 femtoWatt=1×10−15 W). This power level corresponds to a few µV
(microvolt, µ=10−6 ) into a 50 Ω resistance, which is a typical value of input
resistance for a receiver. This signal level must be increased so that it can be
demodulated, and further increased so that it can be heard. The device that
performs this function is called an amplifier.
Amplifiers relate the signal at their input and their output by a gain. We
are usually interested in two types of gain, voltage gain and power gain. We
1.8. MIXERS 9
Vi A Vo
dB −3 0 3 6 7 10 13 16 20 30 40
A 0.70 1.0 1.4 2.0 2.2 3.2 4.5 6.3 10 32 100
G 0.50 1.0 2.0 4.0 5.0 10 20 40 100 1000 10000
where the logarithm function is with respect to base 10. The coefficient is 10
for power gain and 20 for voltage or current gain. With this definition, both
a voltage gain and the corresponding power gain yield the same value in dB.
For example, if a peak voltage of V1 appears across a resistor R, then the peak
current through R is V1 /R, and the average power delivered to R is V12 /2R.
Now, if this voltage is amplified two folds and applied across the same resistor,
then there is a voltage gain of A = 2 and a power gain of G = 4. In decibels,
the value of both AdB and GdB is 6 dB. √ Also, note that 3 dB corresponds to a
power gain of 2 and a voltage gain of 2.
Decibel notation can also used to define absolute levels. For example, 0.5 mil-
liwatt of power is expressed in decibels as −3 dBm. Here, “m” denotes that
this value is relative to 1 milliwatt. Similarly, 20 Watts can be expressed as
43 dBm. Another way of writing absolute levels in decibels is to directly write
what it is relative to. For example, we can write 32 µV as “30 dB re µV”. Some
easy-to-remember approximate dB values are given in Table 1.3.
1.8 Mixers
We frequently want to shift the frequency of the information carrying sinusoid.
For transmission purposes, we want to increase the frequency. The process of
1.9. FILTERS 10
RF IF AB cos (ω +ω )t + cos (ω -ω )t
A cos (ω1t)
2
( 1 2 1 2 )
LO
B cos (ω2t)
1.9 Filters
We employ filters to eliminate unwanted components of a signal and keep the
components we like. Most commonly used filters are classified as low-pass, high-
pass, and band-pass types.
As the name implies, a low-pass filter (LPF) allows the signals below a spe-
cific frequency to pass through the filter and attenuates (decrease their ampli-
tude) the signals of higher frequency. This threshold frequency is called cut-off
or corner frequency. This behavior of an LPF is demonstrated in Fig. 1.10: The
upper schematic demonstrates the removal of the sinusoidal component above
the corner frequency. In the lower schematic, where the input has two sinusoids
at two different frequencies, the higher frequency sinusoidal signal is rejected at
the output.
A high-pass filter (HPF) has the opposite function: It passes signals of high
frequency, while stopping low-frequency signals. (See Fig. 1.11.)
A band-pass-filter (BPF) passes signals within a specified frequency range.
It eliminates signals outside this band. Suppose we wish to keep the fundamental
component of the square wave of Fig. 1.4 at ω and eliminate the rest of the
components. For this purpose, we can use a band-pass-filter centered at ω. The
1.10. TRANSMITTER AND RECEIVER 11
LPF
LPF
upper diagram in Fig. 1.12 demonstrates this action. The bandwidth is of the
filter must be sufficiently small to attenuate the nearest harmonics. Referring to
Fig. 1.5, the amplitude of the fundamental component is 4/π ≈ 1.27. Conversely,
if we use a band-pass-filter centered at 5ω, we pick the 5th harmonic of the square
wave with a peak amplitude of 4/(5π) ≈ 0.25 as shown in the lower diagram of
Fig. 1.12.
The filtering effect is not abrupt, but it is gradual. The signal components
and noise beyond cut-off frequency are not entirely eliminated, but attenuated
more and more as their frequencies are further away from cut-off. A detailed
discussion of filters can be found in Chap. 5.
HPF
HPF
original signal. The signal is fed to a speaker (with a size comparable to the
sound wavelength) to generate the sound waves. The sound waves are probably
heard by a human ear, whose size is also comparable to wavelength.
Since our transmitter is not the only one in the area, the receiver antenna
receives many signals from different sources. Although it is not shown in the
simplified diagram, filtering the antenna input signal that rejects all unwanted
signals is necessary for a clean reception.
1.11 TRC-11
Transceivers are wireless transmitters (TX) and receivers (RX) combined in
a single instrument. TRC-11 (see Fig. 1.15) is a transceiver operating in the
28 MHz amateur band, where a license for transmission is not necessary if the
output power is kept below a specific limit. Hence TRC-11 transmitter output
power is intentionally kept low, not to violate local electromagnetic radiation
regulations. On the other hand, the receiver sensitivity is very good, providing
communication over some distance.
TRC-11 utilizes the superheterodyne principle, which is used by most modern
radio receivers (for example, those in mobile phones) today. Superheterodyne
receiver systems use a frequency down-conversion mechanism of a mixer driven
by an oscillator: The incoming AM modulated signal is mixed with a constant
amplitude sinusoidal wave of a different frequency generated by a local oscilla-
tor. The mixer output is an AM signal at a lower and fixed frequency known
as intermediate frequency (IF), where the signal is more easily amplified in a
narrow-band amplifier chain. If a different frequency input signal is desired, it
is sufficient to change the frequency of the local oscillator, while the frequency
of narrow-band IF amplifier remains unchanged. The same principle can be
used to up-convert a low frequency AM signal to its higher frequency version for
transmission purposes. More discussion on the superheterodyne principle can
be found in p. 275.
1.11. TRC-11 13
2
1.27 cos (ωt)
BPF
0
A square wave
at ω ω
2
BPF 0.25 cos (5ωt)
0
A square wave
at ω
5ω
Amp Modulator
In its transmitter, TRC-11 does not use the up-conversion method. Instead,
the transmission frequency is generated directly by an oscillator. The signal
generated at the transmission frequency is amplitude-modulated and amplifier
to be fed to the antenna.
A block diagram of TRC-11 is shown in Fig. 1.16. A low frequency audio
input signal to the transmitter is amplified by an audio amplifier. The ampli-
fied signal is then used to amplitude modulate a 27.00 MHz sinusoidal signal
generated by an oscillator. This 27.00 MHz AM signal is amplified by a radio-
frequency (RF) amplifier and then fed to the antenna for transmission into air.
A switch (T/R switch) is used to select the transmit or receive mode for the
antenna.
A small amplitude 27.00 MHz AM signal picked by the antenna is fed to
mixer that acts like a down-converter. The mixer uses a 12.00 MHz local os-
cillator as the transmitter. The mixer’s output has two signals: The difference
frequency at 27.00-12.00=15.00 MHz and the sum frequency at 27.00+12.00=
=39.00 MHz. 15.00 MHz signal is the desired IF frequency: It is filtered by a
15.00 MHz narrow-band crystal band-pass-filter, providing the good selectivity
of the receiver. The resulting signal is amplified by a high-gain IF amplifier
chain. The amplitude demodulator block strips the AM signal of its carrier and
generates the original audio signal. This signal is then fed to a loudspeaker (or
1.11. TRC-11 14
Amp Demodulator
earphone) amplifier to drive the loudspeaker (or earphone) generating the audio
signal. The output of the amplitude demodulator is also used for the automatic
gain control circuit, which reduces the gain of the IF amplifier chain and hence
prevents a saturation if the input signal is too strong.
If the antenna also picks a neighboring signal at 26.95 MHz, this signal
will be down-converted to 26.95-12.00=14.95 MHz using a local oscillator of
12.00 MHz. Since the IF filter is strictly at 15.00 MHz, the 14.95 MHz signal
will be rejected. Therefore, very selective TRC-11 receiver only amplifies signals
at 27.00 MHz.
Although it is possible to build a transceiver with modern complex integrated
circuits in a much smaller area, for the purpose of learning and ease of soldering,
TRC-11 is intentionally built from many discrete, inexpensive, —and some old-
fashioned— components.
In the following chapters, we will study all blocks of TRC-11 starting from
the voltage regulator unit. More information about these subjects can be found
1.11. TRC-11 15
TRANSMITTER
27 MHz
Audio Oscillator 27 MHz AM
In
12 MHz +12 V
Local
Oscillator
Automatic
Gain Earphone
Control
27 MHz BPF 15 MHz BPF
LO
RF IF IF Amplitude EP/LS
Amp Demodulator Amp
Mixer Crystal filter
RECEIVER
Loudspeaker
in a book by Nahin [2] or in any yearly edition of the Handbook for Radio
Communications published by the American Radio Relay League [3].
1.12. PROBLEMS 16
1.12 Problems
1. Find the wavelengths of sound waves in air at frequencies of 20 Hz and
20 kHz. (20 Hz and 20 kHz are accepted to be lowest and highest audible
sound frequencies for most humans.)
8. Construct a square wave with three and five components and calculate the
mean square error, using a computer tool of your choice, a spreadsheet,
MATLAB, etc. Mean-square error, M SE, between two periodic functions,
f (t) and g(t) is given by
s
1 T
Z
M SE = (f (t) − g(t))2 dt
T 0
CIRCUIT THEORY
PRIMER
2.1 Electrons
Atoms of some materials, notably metals or acids, allow the movement of the
electrons readily. We classify such materials as conductors. Gold and aluminum
are excellent conductors, while iron and lead are not. Some materials, such as
niobium-titanium alloy, exhibit perfect conductivity, known as superconductivity
below a critical and relatively low temperature. On the other hand, atoms of
insulators do not allow the electrons to move at all. Dry wood, porcelain, quartz,
and rubber are good insulators. Materials, such as germanium and silicon,
whose electron conductivity falls midway between good conductors and good
insulators, are known as semiconductors. The addition of a small percentage
of foreign atoms into semiconductors, which is known as doping, changes the
conductivity properties dramatically. Semiconductor devices [6] provide the
enabling technology of the information age.
Electronics are all about controlling electrons. To quantify the movement of
electrons in a circuit, we use a number of terms:
Charge represented by Q is used to measure the number of electrons.
* ARRL Handbook of Radio Communications [3] has comprehensive information on elec-
tronic components
It has the units of coulombs (unit symbol C), named after the French
physicist Charles-Augustin de Coulomb (1736–1806). He is known for
developing the Coulomb’s law. Since an electron has a negative charge,
6.241×1018 electrons make up a charge of −1 C. Equivalently, the charge of
one electron is −1.602 × 10−19 C. Positive charges may also exist. An elec-
trolyte may have positively charged ions in addition to negatively charged
ions. For example, salty water has positive Na+ and negative Cl− ions.
Current measures the flow rate of charged particles, represented by the
letter I. If 1 C of charge moves in one second, it is called one am-
pere (unit symbol A), named after French physicist André-Marie Ampère
(1775–1836) known for developing Ampere’s law. Current is a directional
quantity. The current direction is the same as the flow direction of posi-
tive charges. Hence, 6.241×1018 electrons moving left create a current of
1 A towards the right. The current in a circuit can be measured using an
ammeter.
Voltage quantifies the electrical potential difference between two points
in a circuit. It measures the desire for charges to move from one place
to another. It has the unit volt (unit symbol V), named after Italian
physicist Alessandro Volta (1745–1827) , who invented the first chemical
battery. 1 V of voltage can deliver 1 J (joule) of energy to 1 C of charge
(E = QV ). The potential difference between the two points is measured
using a voltmeter.
Resistance defines the degree to which a conductor opposes the electric
current through it. The unit of resistance is one ohm (unit symbol is
Greek letter capital omega Ω), named after German physicist Georg Simon
Ohm (1789–1854). A good conductor like a copper wire has a very low
resistance; the electrons flow freely through it. Water is a relatively poor
conductor of current, so it has a higher resistance. Insulators like glass or
ceramics have very high resistance, with negligible current through them.
Resistivity represented by the Greek letter ρ is an intrinsic property of a
material that quantifies how strongly that material opposes the flow of
current. The resistivity of some common materials is listed in Table 2.1.
The unit of resistivity is Ω-cm and it defines the resistance between the
opposing faces of one cubic centimeter of the material.
The resistance of a material with a resistivity of ρ, a cross-sectional area
of A and a length of l is given by
ρl
R= (2.1)
A
Electrical components called resistors have a predefined level of resistance,
represented by the symbol R. Their typical values range from 1 mΩ to
1 GΩ (G means 109 ). The resistance of a component can be measured
using an ohmmeter.
2A
+ + +
+ +
8V 2V 2A 1A 5V 2A 5V
3 cos(ωt) 3A -
Figure 2.3: (a) Not allowed: a short-circuited voltage source, (b) not allowed:
two voltage sources with different values in parallel, (c) not allowed: an open-
circuited current source, (d) not allowed: two current sources with different
values in series (e) allowed: a current source in parallel with a voltage source.
with the directions of voltages and currents as shown in Figs. 2.1 and 2.2. If the
value of p(t) is positive, power is delivered by the source. On the other hand, if
2.2. ENERGY SOURCES 22
Power is measured in units of watts (W), named after Scottish scientist James
Watt (1736–1819). For example, if a light bulb in a 3 V flashlight draws 100 mA,
the power delivered by the batteries is P = 3 × 0.1 = 0.3 W.
Energy is the work done by an electrical source in given time duration. It is
defined as
Z T
E(T ) = p(t)dt (2.4)
0
Energy is measured in joules (J), named after English physicist James Prescott
Joule (1818–1889). For example, a 12 V car battery of 80 Ampere-hours capacity
stores an energy of E = 12 × 80 × 3600 = 3.4 × 106 = 3400 kJ, while a Ni-Cd
battery of 1.2 V with a 600 mAh capacity stores energy of E = 1.2×0.6×3600 =
2.6 kJ.
Suppose the voltage across an element, v(t), is given by a sinusoid at the
radial frequency of ω as
v(t) = V1 cos(ωt + θv ) (2.5)
and the current, i(t), through it is given similarly by
i(t) = I1 cos(ωt + θi ) (2.6)
where θv and θi are the phases of the voltage and current, respectively. We can
calculate the instantaneous power, p(t), delivered to it as
p(t) = v(t)i(t) = V1 I1 cos(ωt + θv )cos(ωt + θi ) (2.7)
or
V 1 I1 V 1 I1
p(t) = cos(θv − θi ) + cos(2ωt + θv + θi ) (2.8)
2 2
In the case of a resistor, the current and voltage have the same phase (θv =
θi ), and hence we can write the power delivered to a resistor as
V 1 I1 V 1 I1
p(t) = + cos(2ωt + 2θv ) (2.9)
2 2
We see that the phase difference between the voltage and the current in an
element or a branch of a circuit is critical and must be carefully controlled in
many aspects of electronics.
The average power, P , is the average value of p(t) in Eq. 2.8 integrated over
one cycle:
1 T
Z
V1 I1
P = p(t)dt = cos(θv − θi ) (2.10)
T 0 2
For a resistive load, the voltage and current phases are the same. Hence we
have θv = θi , and the average power is
V 1 I1
P = (2.11)
2
2.2. ENERGY SOURCES 23
We note that if the element is such that the phase difference between the
voltage across and current through is θv − θi = 90o , P is zero. Inductors and
capacitors are such elements with no power dissipation.
If the voltage across and current through an electrical element are sinusoidal
with
2π 2π
v(t) = V1 cos t + θv and i(t) = I1 cos t + θi (2.13)
T T
where V1 and I1 are the peak amplitudes, and T is the period. The rms value
of v(t), Vrms , is found after integration operation as
s s
1 T 2 V12 T
Z Z
2π V1
Vrms = v (t) dt = cos2 t + θv dt = √ (2.14)
T 0 T 0 T 2
where T is the period o the sine wave. Similarly, rms value of a sine wave
current, i(t), is
s s
1 T 2 I12 T
Z Z
2π I1
Irms = i (t) dt = cos 2 t + θi dt = √ (2.15)
T 0 T 0 T 2
We note that an AC voltmeter measures the rms (not the peak) value of the
voltage across its terminals, assuming that the voltage is sinusoidal. Similarly,
an AC ammeter measures the rms current √ flowing through it. Since the ratio of
the peak value and to the rms value is 2 for √ a sinusoidal signal, a 220 Vrms
sinusoidal line voltage has a peak value of 220 2=311 V. Using Eq. 2.16, we
deduce that a P =60 W light bulb operating at the line voltage of 220 Vrms
draws a current of 60/220 = 0.27 Arms , since θv = θi for the light bulb.
Example 1
The voltage across a 10 Ω resistor, vR (t), is triangular with a period T as shown
in Fig. 2.4. Find the average power, Pav , dissipated in the resistor.
2.2. ENERGY SOURCES 24
vR(t)
2
T/2
T/4 T
-2
Solution
Due to symmetry, the rms voltage, VRrms , of the resistor can be found by
integrating over only a quarter cycle (between 0 and T /4):
s s s
2 T /4
1 T
Z T /4
4 64t3
Z
4 2t 2
VRrms = vR (t)2 (t) dt = dt = 2
=√
T 0 T 0 T /4 T 3T 0 3
√
(The rms value of a triangular waveform is 1/ 3 times its peak value.) Hence
the average power dissipated in the resistor is
2
VRrms 22 1
Pav = = = 0.133 W.
R 3 10
RS RS1 RS2
+ Is RP + + I1 I2
RP1 RP2
VS V1 V2
Figure 2.5: (a) Equivalent circuit of a real-life voltage source with a voltage
VS and an internal resistance of RS , (b) equivalent circuit of a real-life current
source with a current of IS and a parallel resistance of RP , (c) an allowed circuit:
two real voltage sources of different values in parallel, (d) an allowed circuit: two
real current sources of different values in series.
there is no current drawn from the battery, the voltage across the terminals is
2.2. ENERGY SOURCES 25
Vo . When a load resistance is connected to this battery, the voltage across the
battery terminals is no longer Vo , since there is a voltage drop across RS .
A real-life current source is given in Fig. 2.5(b). It has a parallel resistance
of RP . It is possible to leave it open-circuited without causing a contradiction
because the current Io can go through the resistor RP .
Fig. 2.5(c) and (d) show allowed circuits since the connections do not create
a contradiction due to the presence of resistors.
As far as phase voltage is concerned, 220 Vrms refers to the voltage difference
between any one of the phase voltages and neutral. On the other hand, the
potential difference between any two phases, which is called line-to-line voltage,
e.g., between v1 (t) and v2 (t), is
√
∆v(t) = v1 (t) − v2 (2) = 3Vp sin(ωt − 30o ) (2.19)
√
The potential difference between the phases is, therefore, 3=1.73 times larger
than any one of phase voltages with respect to neutral. The line-to-line voltage
level is 381 Vrms for a phase voltage of 220 Vrms . The last step-down from MV
to low voltage (LV) is depicted in Fig. 2.6. Note that there is no neutral for
When the energy is carried on three phases only, the nominal rms line volt-
ages refer to the potential between the phases. 34.5 kVrms , for example, is the
line voltage in MV lines.
A typical MV to LV transformer configuration is given in Fig. 2.6. The three-
phase line-to-line voltage of 34.5 kV MV is connected to the primary windings
of a three-phase transformer, which is connected in a ∆ configuration. The
secondary terminals are LV terminals, and three windings are now configured
in a Y form. In other words, one terminal of each of the secondary windings
is connected to the earth, while there is no earth connection on the primary.
The voltage transformation ratio in these transformers is always stated as the
ratio of line-to-line voltages (i.e., the potential difference between the phases) of
primary and secondary windings. However, the physical turns ratio of primary
and secondary windings correspond to 34.5 kV to 220 V.
Three 220 V live lines, neutral and earth are distributed in the buildings
through a few distribution panels. Precautions against excessive current are
taken at each panel. This reduces the fire risk in the building and is not helpful
to avoid electric shock. One can get electric shock either by touching both live
and neutral simultaneously, or by touching live while having contact with the
ground.
Building floors have a connection to a ground reference, although there may
be some resistance in between. Therefore if one touches the line while standing
on the floor, e.g., with shoes with natural soles (not an isolating sole like rubber),
he/she gets a shock. It is likely that there is an extra precaution at the last
panel, where a residual current device (RCD) is fitted. This device monitors the
difference between the line and neutral currents, and when it exceeds 30 mA, it
breaks the circuit. This decreases the severity of the shock.
This law is named after German physicist Gustav Robert Kirchhoff (1824–1887).
Fig. 2.7 shows two example circuits. The reference directions for voltages are
assigned arbitrarily by the positions of the + signs. We note that the individual
voltages can be positive or negative depending on the direction assignment. For
2.3. KIRCHHOFF’S CIRCUIT LAWS 28
While going clockwise around the loop, we use a plus sign in KVL equa-
tion if the first-encountered sign of a component voltage is plus, a negative
sign otherwise.
Using the same convention for the second circuit in (b), the KVL equations
for the three loops are:
−VS + V1 + V2 − V3 = 0
−VS + V1 − V4 + V5 − V3 = 0
−V2 − V4 + V5 = 0
We do not have three independent equations. The third equation can be ob-
tained by subtracting the first one from the second.
From the conservation of charge, Kirchhoff’s current law states that the
sum of currents flowing into any node (A node is a point in the circuit
where more than two elements are connected together) should be equal to
the sum of currents leaving that node. For a node with N branches, it can
be written as
XN
In = 0 (2.21)
n=1
where the branch currents flowing into the node have a negative sign, and
the branch currents leaving the node have a positive sign.
Fig. 2.8 shows two example circuits. In the first circuit of (a), we choose
directions for currents arbitrarily. I2 can be found from Ohm’s Law as I2 =
To simplify the notation, we use 1000 Ω=1KiloΩ=1K and 1,000,000 Ω=1MegaΩ=1 M.
R
- V + +V +
+ R + 1
+ V4
+
VV VI VS V2
-
- V3 V5
+ +
(a) (b)
Figure 2.7: (a) An example circuit with three components and one loop, (b) an
example circuit with six components and three loops
2.4. RESISTORS AND OHM’S LAW 29
V = RI (2.22)
where R is the resistance of the resistor. In a given resistor, more current flows,
if more voltage is applied.
In the water flow analogy illustrated in Fig. 2.10(a), more water will flow
through a pipe if the pressure across the pipe increases. A variable resistor is
analogous to a water tap (Fig. 2.10(b)).
Example 2
Let us determine the resistance of a 1 mm diameter copper wire of 100 m length.
Using Eq. 2.1 and Table 2.1 we write
ρl (1.68 · 10−8 Ω-m)(100 m)
R= = = 2.14 Ω (2.23)
A π(0.5 · 10−3 )2 m2
I1 I3 I1 I4 2µΑ
A B
I2 I2 I3
+ +
5V 1K 3mA 1M 6V 1.5V
5µΑ +
2M
(a) (b)
Figure 2.8: (a) An example circuit with three components, (b) an example
circuit with six components
2.4. RESISTORS AND OHM’S LAW 30
r[ht]
I
I
+
slope=1/R
R V
- 0 V
pipe tap
water water
(a) (b)
Figure 2.10: (a) Resistor analogy: a water pipe with a constriction, (b) Variable
resistor analogy: a water tap
v 2 (t)
p(t) = v(t) i(t) = = i2 (t) R (2.24)
R
If the applied signals are periodic, the average power, P , can be found using an
average over a period T :
1 T 1 1 T 2 1 T 2
Z Z Z
P = v(t)i(t)dt = v (t)dt = R i (t)dt (2.25)
T 0 RT 0 T 0
From Eq. 2.12, we have
Z T Z T
2 1 1
Vrms = v 2 (t)dt and Irms
2
= i2 (t)dt (2.26)
T 0 T 0
and hence we can simplify the expressions in Eq. 2.25. The average power
dissipated on R is
V2
P = rms = Irms2
R (2.27)
R
in unit of watts.
Real-life resistors have power dissipation limit. If this limit is exceeded, a
resistor may get destroyed.
2.4. RESISTORS AND OHM’S LAW 31
Example 3
Consider the circuit of Fig. 2.11. Find the power delivered/dissipated by the
sources and the resistor. Check that the total power delivered is equal to total
power dissipated. Note that the reference directions for resistor and for sources
are different.
IV
+
I R IR V
2A 50Ω 12V
Solution
The voltage across the resistor is determined by the voltage source. The current
through the resistor is IR = V /R = 12/50 = 0.24 A. Hence the power dissipation
in the resistor is
12
PR = V I R = = 2.88 W
0.24
The voltage across the current source is also determined by the voltage source.
The power delivered by the current source is
PI = V I = 12 × 2 = 24 W
The current through the voltage source is determined by KCL: IV = IR − I =
0.24 − 2 = −1.76 A. The power delivered by the voltage source is
PV = V IV = 12 × (−1.76) = −21.12 W
Since PV is negative, the power is not delivered but rather absorbed by the
voltage source. The total power delivered is equal to the total power dissipated:
PI + PV = 24 − 21.12 = 2.88 = PR
The power delivered by the current source is partly dissipated in the resistor
and the remaining part is absorbed by the voltage source.
2. significant
figure Tolerance
1. significant Exponent
figure
Color Significant figure Multiplier Tolerance
Black 0 ×100
Brown 1 ×101 ± 1%
Red 2 ×102 ± 2%
Orange 3 ×103
Yellow 4 ×104
Green 5 ×105 ± 0.5%
Blue 6 ×106
Violet 7 ×107
Gray 8 ×108
White 9 ×109
Gold ×10−1 ± 5%
Silver ×10−2 ± 10%
Table 2.2: Resistor color codes. The resistor shown above has a 4-band code:
15 kΩ with 10% tolerance.
10, 12, 15, 18, 22, 27, 33, 39, 47, 56, 68, and 82.
On the other hand, the resistors with 5% tolerance can be found in the following
values:
10, 11, 12, 13, 15, 16, 18, 20, 22, 24, 27, 30, 33, 36, 39, 43, 47, 51, 56, 62, 68,
75, 82 and 91.
Figure 2.12: Leaded resistors with 1/4W, 1W, and 11W power ratings.
Figure 2.13: The symbol of a potentiometer (top left), of a trimpot (top right).
Photos of a potentiometer (bottom left) and trimpots of different types. Multi-
turn trimpots have high degrees of accuracy (second from the bottom right).
V = V1 + V2 = R1 I + R2 I = (R1 + R2 )I (2.28)
R1 I R2 R1 I R2 Rn
...
+V - +V - +V - +V - + V -
1 2 1 2 n
+ V - + V -
(a) (b)
I1 R1
I1 R1
I I I2 R2
I2 R2
...
...
+ V - In Rn
+ V -
(a) (b)
where we used Ohm’s law for each resistor. Therefore, in a series connection of
two resistors, the total resistance is equal to the sum of the resistances:
R = R1 + R2 (2.29)
Series connected resistors are used frequently as a voltage divider. The voltage,
V1 , across R1 can be written in terms of V as
R1
V1 = V (2.30)
R1 + R2
If there are n resistors in series as in Fig. 2.14(b), the voltage across the resistor
R1 can be found from
R1
V1 = V (2.31)
R1 + R2 + . . . + Rn
Since the voltage dividers are very common, it is worth learning the formula
above.
When n resistors are connected in series, the total resistance can be found
easily:
R = R1 + R2 + · · · + Rn (2.32)
R2
I1 = I (2.35)
R1 + R2
Note that unlike Eq. 2.30, the resistance in the numerator refers to the resistor
in the other branch.
When n resistors are connected in parallel as shown in Fig. 2.15(b), the total
resistance can be found from
−1
1 1 1
R = R1 ∥ R2 ∥ · · · ∥ Rn = + + ··· + (2.36)
R1 R2 Rn
R2 ∥ R3 ∥ . . . ∥ Rn
I1 = I (2.37)
R1 + (R2 ∥ R3 ∥ . . . ∥ Rn )
where the numerator contains the parallel combination of all resistors in the
other branches.
R2 c R4 d
RT1 R1 R3 RT2 R5 R6
(a) (b)
R1 (R2 + R3 )
RT 1 = R1 ∥ (R2 + R3 ) =
R1 + R2 + R3
In Fig. 2.16(b), two resistors are first connected in parallel, then connected in
series with another resistor. The equivalent resistor RT 2 can be found similarly
as
R5 R6
RT 2 = R4 + (R5 ∥ R6 ) = R4 +
R5 + R6
2.5. ANALYSIS OF ELECTRICAL CIRCUITS: NODAL ANALYSIS 36
1. Select a common node with the most possible branches so that all
other node voltages are defined with respect to this node. Call this
node the ground node.
2. Define the voltage difference between all other nodes and the ground
node as the unknown node voltages.
3. Write down the KCL at each node, expressing the branch currents
in terms of node voltages and sources. (Not to get confused, write
the branch currents always as leaving the node except when there are
current sources.) If there is a voltage source between two nodes, write
down the KVL between those two nodes.
4. Solve the equations obtained in step 3 simultaneously.
5. Find all branch currents and voltages in terms of node voltages.
Example 4
Consider the circuit in Fig. 2.17. Let us analyze this circuit to find all element
VA R2 VB
A
I1 R1 R3
Example 5
Consider the circuit of Fig. 2.18 containing three sources and five resistors. We
apply the nodal analysis procedure:
1. Assign the ground symbol to the bottom node.
2. We have three nodes in this example. We assign Va , Vb , and Vc as node
voltages.
3. We note that Va is already known as Va = 8V. Therefore, we need to write
only two equations. Write KCL at nodes b and c as
Vb − 8 Vb Vb − Vc
+ + − 6 mA = 0
3K 5K 1K
Vc − 8 Vc Vc − Vb
+ + + 10 mA = 0
2K 4K 1K
We note that (V, kΩ, and mA) is a consistent unit set.
2.5. ANALYSIS OF ELECTRICAL CIRCUITS: NODAL ANALYSIS 38
2K
Va 3K Vb 1K Vc
+
8V 6m 5K 4K 10m
Example 6
Vx
2.2M 1.2M
+
Vy 3.3M Vz
10V
1M
0.82M
5µΑ
Note that we cannot use the current divider or voltage divider equation to solve
this circuit.
2.6 Capacitors
Capacitors are built from two conductor plates separated by a thin insulator, as
shown in Fig. 2.20. When a voltage of V is applied across the plates, electrical
charges +Q and −Q accumulates at the plates. The charges remain there even
after the voltage is removed. Hence they act like charge storage devices. In this
respect, they resemble voltage sources. However, they cannot supply constant
voltage for a long time since they have a finite reservoir of charge. The ratio
+ +Q
+ + + + + + + + +
Conductor plates
V d
- - - - - - - - -
of area A
- -Q
Figure 2.20: Structure of a capacitor built with two conducting plates separated
by an insulator.
ϵo ϵr A
C= (2.39)
d
where ϵo = 8.85 × 10−12 F/m is the permittivity of free space and ϵr is a unitless
quantity showing the relative permittivity of the insulator used between the
plates. Table 2.6 lists the relative permittivities of some materials. The symbols
of non-polarized, polarized, and variable capacitors are depicted in Fig. 2.21.
2.6. CAPACITORS 40
Example 7
Find the capacitance of a capacitor built by a paper dielectric with conductor
planes of 3 cm×10 cm. The paper has a thickness of 0.1 mm.
i(t) i(t)
+ + +
v(t) v(t)
- -
Figure 2.21: Symbols for capacitors: (a) Non-polarized, (b) polarized (c) vari-
able
Rubber
diaphragm
water
i(t)
C + + C
I=1 mA v(t) v(t)=
10nF - 3 sin(ωt) 1 µF
v(0)=2 ω=2π104
(a) (b)
Figure 2.23: (a) A current source driving a capacitor, (b) an AC voltage source
across a capacitor.
more water to increase the pressure. On the other hand, if the diaphragm is
made stiffer, the capacitance of the tank is reduced.
If we let a current, i(t), of arbitrary time waveform, pass through a capacitor,
the amount of charge accumulated on the capacitor within a time interval, 0 to
t1 , is given as Zt1
∆Q = i(t)dt (2.41)
t=0
where ξ is the dummy variable of integration. Q(0) refers to the initial charge
on the capacitor at the time instant t = 0. Using Eq. 2.38, we can relate the
voltage across a capacitor and the current through it
Z t
1
v(t) = v(0) + i(ξ)dξ (2.43)
C ξ=0
Example 8
Referring to Fig. 2.23(a) and using Eq. 2.43, the voltage of the capacitor driven
by the current source is given by
1 1
v(t) = v(0) + It = 2 + 10−3 t = 2 + 105 t (2.45)
C 10 × 10−9
2.6. CAPACITORS 42
The current in the capacitor of Fig. 2.23(b) can be found using Eq. 2.44:
d
i(t) = C (3 sin(ωt)) = 3(1 · 10−6 )(2π104 ) cos(ωt) = 0.19 cos(ωt) (2.46)
dt
There are two major types of capacitors. The first type is non-polar, i.e., the
voltage can both be positive and negative. Most of the capacitors of smaller than
0.5 µF (µ=micro=10−6 )¶ are of this type. However, as the capacitance values
become large, it is less costly to use capacitors, which have polarity preferences,
like electrolytic or tantalum capacitors. Fig. 2.24 is a photo of leaded tantalum
and electrolytic capacitors with different voltage ratings. For these capacitors
the voltage must always remain in the same polarity indicated on the capacitor.
For values less than 1 nF, variable capacitors are also available (see Fig. 2.25).
C1 Q1
V1 V2
+ - + -
+ V -
C1 Q C2 Q
C2 Q2 + V -
(a) (b)
C = C1 + C2 (2.48)
In the water tank analogy, obviously, the tank capacities are added when
two tanks are connected in parallel as shown in Fig. 2.28(a).
water
water water
water
(a) (b)
Figure 2.28: Water tanks with the diaphragms (a) in parallel, (b) in series
2.7 RC Circuits
When a resistor, R, is connected to a charged capacitor, C, in parallel, as
in Fig. 2.29(a), the circuit voltages become a function of time. Assume that
initially capacitor is charged to V0 volts (it has Q = CV0 coulombs stored
charge). At t=0 we connect the resistor, R. Using KCL and Ohm’s law, we
write
dv(t) v(t)
i(t) = C =− (2.53)
dt R
Since the current i(t) is leaving the resistor in the opposite direction to the voltage v(t),
we must use a negative sign for v(t)/R.
2.7. RC CIRCUITS 45
V0
i(t) +
C v(t) R v(t)
-
t
(a) (b)
or
dv(t) v(t)
+ =0 (2.54)
dt RC
This equation is called a first-order differential equation. Its solution for
t ≥ 0 is
v(t) = V0 e−t/RC for t ≥ 0. (2.55)
v(t) is plotted in Fig. 2.29(b). The above expression tells us that as soon as the
resistor is connected, the capacitor voltage starts decreasing, i.e., it discharges
on R. The speed with which discharge occurs is determined by τ = RC. τ is
called time constant and has units of time (1Ω× 1F=1sec). The current flowing
in the capacitor is
dv(t) d V0
i(t) = C = C V0 e−t/RC = − e−t/RC for t ≥ 0. (2.56)
dt dt R
The value of the current is found negative, indicating that the current flows
in the direction opposite to the one shown in the figure.
Note that we used a sign convention while writing Eq. 2.53: The sign of the
voltage on an element must be chosen positive if the positive terminal is the one
where the current enters the element. In this circuit, the current is chosen in
the direction entering to C at its positive terminal (top) (as in Fig. 2.21), thus
the sign of the capacitance equation of Eq. 2.53 must be positive. On the other
hand, the current leaves the positive terminal for the resistor (unlike the current
in Fig. 2.9); hence we must use a negative sign for Ohm’s law.
Usually, we do not know the actual current directions and voltage polarities
when we start the analysis of a circuit. We assign directions and/or polarities
arbitrarily and start the analysis. We must carefully stick to the above con-
vention when writing down the equations of the elements or KVL and KCL
equations. Otherwise, our results are not correct.
The magnitude of the current in the above circuit is at its maximum, V0 /R,
initially, and decreases towards zero as time passes. This is expected as the
voltage across the capacitor similarly decreases.
2.8. ANALYSIS OF FIRST-ORDER RC CIRCUITS 46
hydraulic capacitance
hydraulic resistance
Figure 2.30: Water flow analogy of RC circuit: water tank feeding a hydraulic
resistor
ginning, the diaphragm is highly stretched, and the pressure is the greatest.
At that time, the water flow will be the fastest. As the water flows, the di-
aphragm relaxes and the pressure reduces. The flow rate approaches zero when
the diaphragm relaxes fully. The energy stored in the stretched diaphragm is
dissipated in the hydraulic resistor.
1. Kill the sources: Place a short-circuit for the voltage sources, and
remove (or open-circuit) the current sources. Find the equivalent
resistance, Req , across the capacitor.
2. Write the time constant as τ = Req C
Note that when we substitute t=0, we get v(0) = vi and i(0) = ii , consistent
with step 3. Likewise, as t → ∞, v(t) approaches to vf , and i(t) approaches to
if , consistent with step 4.
If a circuit contains two or more independent capacitors, the circuit is no
longer a first-order circuit. Obtaining the time-domain solution of such circuits
is more difficult, and we do not deal with such circuits. A circuit simulator can
be used for that purpose. See page 300 for a tutorial on time-domain solutions
of circuits using LTSpice.
Example 9
vC(0)=-2V
+ vR - + vR - + vR - + vR -
+ 2K 2K + 2K + 2K
+ +
5V 3µF vC 3µF 5V 2V 5V
+
Consider the simple circuit of Fig. 2.31(a) with a single capacitor. The
initial value of the capacitor voltage is given at t = 0. Suppose we would like
to determine the voltage across the resistor, vR (t) in the polarity shown. Apply
the procedure:
1. Kill the voltage source: Place a short-circuit in its place as depicted in
Fig. 2.31(b). The equivalent resistance across the capacitor is Req =2K
vR (t) = 0 + (7 − 0)e−t/6 m
= 7e−t/6 m
2.8. ANALYSIS OF FIRST-ORDER RC CIRCUITS 48
i1 i1
vC(0)=2V
+ 4K 4K
+ +
10V 6µ vC 2K 3m 6µ vC 2K
(a) (b)
i1 i1
vC
+ 4K + 4K
+
10V 2V 2K 3m 10V 2K 3m
(c) (d)
Example 10
Consider the circuit of Fig. 2.32(a) with a single capacitor. The initial value of
the capacitor voltage is given at t = 0. We would like to determine the current
through the 4 kΩ resistor, i1 (t) in the direction shown with two significant
figures. Apply the procedure:
1. Kill the sources: Place a short-circuit for the voltage source and an open-
circuit for the current source as in Fig. 2.32(b). The equivalent resis-
tance across the capacitor is 2 kΩ in parallel with 4 kΩ. From Eq. 2.34:
Req =(2K)∥(4K)=(2K 4K)/(2K+4K)=1.3 kΩ
2. Write the time constant as τ =(1.3 kΩ)(6 µF)=8.0 ms
3. Substitute a voltage source with a value equal to the initial voltage of
the capacitor as in Fig. 2.32(c). From KVL and Ohm’s law we find
i1i =(10−2)/4K=2.0 mA.
4. Substitute an open circuit for the capacitor as in Fig. 2.32(d). Writing
the node equation at vC :
vC − 10 vC
+ − 3 mA = 0
4K 2K
We find vC =22/3=7.3 V. Hence i1f =(10-22/3)/4K=2/3=0.67 mA.
5. Write the solution for the desired current as
2 2
i1 (t) = + (2 − )e−t/8m = 0.67 + 1.3e−t/8m mA
3 3
A MATLAB code to plot i1 (t) is given below. The corresponding graph is shown
in Fig. 2.33 with the tangent to the curve drawn at t=0. Note that the tangent
line intersects the final value line i1f =0.67 mA at t = τ =8 ms.
Example 5
2
i1 (t)
1.8
Tangent at t=0
Asymptote at infinity
1.6
1.4
1.2
i 1 (mA)
0.8
0.6
0.4
0.2
0
0 5 10 15 20 25 30 35 40
t (ms)
Example 11
Consider the circuit of Fig. 2.34(a) with a single capacitor. The initial value of
the capacitor voltage is given at t = 0. We would like to determine the current
through the 4 kΩ resistor, i2 (t), in the direction shown. Apply the procedure:
1. Kill the sources: Place a short-circuit for the voltage source and an open-
circuit for the current source as in Fig. 2.34(b). The equivalent resistance
across the capacitor is:
2K 3K 4K 1K
Req = (2K) ∥ (3K) + (4K) ∥ (1K) = + = 2K
2K + 3K 4K + 1K
vC(0)=-6V
2K 4K 2K 4K
+ vC i2 vC i2
+ +
5V
1µ 1µ
3K 1K 3K 1K
(a) (b)
V1
2K 4K 2K 4K
+ 6V + i2 + i2
5V V2 V3 5V
3K 1K 3K 1K
(c) (d)
3. Substitute a voltage source with a value equal to the initial voltage of the
capacitor as in Fig. 2.34(c). Use nodal analysis to solve the circuit:
V3 : V3 = V2 + 6
V2 −5 V2 V3 −5 V3
V2 : 2K + 3K + 4K + 1K =0
2.9 Inductors
When a current flows through a wire, a magnetic flux is generated around
the wire. Reciprocally, if a conductor is placed in a time-varying magnetic
field, a voltage is induced in it. From the electrical circuits point of view,
this phenomenon introduces the circuit element, the inductor. Inductors are
2.9. INDUCTORS 51
+ i(t)
+ i(t) v1 L1 + i1 i2
-
v(t) + v L1 L2
- v2 L2 -
-
Figure 2.35: (a) Inductor symbol, (b) inductor symbol with a core, (c) inductors
in series, and (d) inductors in parallel.
Z t
di(t) 1
v(t) = L or i(t) = i(0) + v(ξ)dξ (2.58)
dt L 0
where L is the inductance in Henries, i(t) and v(t) are current through, and
voltage across the inductor. i(0) is the current of the inductor at t=0.
We note that voltage is proportional to the time derivative of current in an
inductor. If i(t) is a DC current, its derivative is zero, and hence the voltage
induced across the inductor is zero. Alternatively, if we apply a DC voltage
across an inductor, the current increases linearly (we can only do this for a
short time; otherwise, the current through the inductor can be very large).
♦ TRC-11 has three inductors. They are used to form tuned circuits in
conjunction with capacitors.
2.10. RL CIRCUITS 52
L = L1 + L2 + · · · + Ln (2.59)
whereas for parallel-connected inductors (Fig. 2.35(d)), voltages are the same
and the currents add up:
−1
1 1 1
L= + + ··· + (2.60)
L1 L2 Ln
2.10 RL circuits
Suppose a resistor, R, is connected in parallel with an inductor, L, as in
Fig. 2.38(a). Initially, the inductor current is equal to iL (0) = Ii . Since R
2.10. RL CIRCUITS 53
and L are connected in parallel, they have the same terminal voltage:
diL
v = RiR = L (2.62)
dt
and from KCL we have
iL (t) = −iR (t) (2.63)
Therefore,
diL (t)
L + RiL (t) = 0 (2.64)
dt
This equation is again a differential equation similar to the capacitor discharge
equation. Its solution is also similar:
flywheel
water
hydraulic resistance
beginning, the flywheel with a moment of inertia is turning fast and the flow
rate is the greatest. At that time, the pressure across the hydraulic resistor
is the greatest. As the water flows, the flywheel slows down and the pressure
reduces. The flow rate becomes zero, when the flywheel stops. The energy
stored in the flywheel is dissipated in the hydraulic resistor.
2.11. ANALYSIS OF FIRST-ORDER RL CIRCUITS 54
1. Kill the sources: Place a short-circuit for the voltage sources, and
remove (open-circuit) the current sources. Find the equivalent resis-
tance, Req , across the inductor.
Example 12
i1 i1
iL(0)=5mA
+ 5K 5K
iL
20V 1K 1K
3mH 10mA 3mH
(a) (b)
i1 i1
V1
+ 5K + 5K
20V 1K 20V 1K
5mA 10mA 10mA
(c) (d)
Consider the circuit of Fig. 2.40(a) with a single inductor. The initial value
of the inductor current is given at t = 0. We would like to determine the current
through the 5 kΩ resistor, i1 (t) in the direction shown. Apply the procedure:
1. Kill the sources: Place a short-circuit for the voltage source and an open-
circuit for the current source as in Fig. 2.40(b). The equivalent resistance
2.11. ANALYSIS OF FIRST-ORDER RL CIRCUITS 55
i1 (t) is plotted in Fig. 2.41 where the tangent to the curve at t=0 is shown.
It intersects i1 =4.0 mA line at t = τ = 3.6 ms.
Example 9
6
4
i1 (mA)
0
0 2 4 6 8 10 12 14 16
t (µs)
Example 13
Consider the circuit of Fig. 2.42(a) with a single inductor. We would like to find
the voltage across the inductor. The initial value of the inductor current is not
given. Instead, we know that the voltage source VS is 4 V for t < 0. So, the
inductor current must have reached its final value for t < 0. First, apply the
part of the procedure to find iL (0− ) (it is the final value of the inductor current
when VS =4 V). Since the inductor current must be continuous, we must have
iL (0− ) = iL (0+ ).
2.12. IDEAL TRANSFORMER 56
+ 200 + 200
4V t<0 + iL + iL +
4V
-5V t>0 − VS vL VS vL
8mH
(a) (b)
200 + 200
iL + +
vL -5V VS vL
8mH 20mA
(c) (d)
vL (t) = −9e−t/40µ
vL (t) is plotted in Fig. 2.43. Note that the inductor voltage is discontin-
uous at t=0, while the inductor current is continuous.
Example 10
1
−1
−2
−3
vL (V)
−4
−5
−6
−7
−8
−9
−50 0 50 100 150 200
t (µs)
i1 n1 : n2 i2
+ +
v1 v2
- -
that appear across the primary winding to another pair of amplitudes at the
secondary, and vice versa. The amount of transformation is determined by the
turns ratio n2 /n1 . The relations in an ideal transformer are as follows:
v2 n2 i1 n2
= and = (2.67)
v1 n1 i2 n1
flywheel
water
flywheel
water
2.13.1 Varistors
Varistors are nonlinear resistors made of ceramic-like materials like sintered zinc
oxide or silicon carbide. The I − V characteristics of a varistor are depicted in
Fig. 2.46, together with its symbol. When the voltage across the varistor is
operating I
range +
V
V -
within the operating range, varistor exhibits a very large resistance. When the
voltage increases, the resistance falls rapidly, thus taking most of the excess
current due to overvoltage. Varistors are connected in parallel to the circuits to
be protected.
2.13.2 Thermistors
A thermistor is a resistor whose resistance depends on the temperature. There
are two kinds: Negative-temperature coefficient (NTC), resistance decreases as
temperature rises; positive-temperature-coefficient (PTC), resistance increases
as temperature rises.
PTC thermistor
We consider a PTC thermistor as a resettable fuse. PTC is in low-resistance
state at room temperature, but its resistance increases abruptly with increas-
ing temperature beyond a specified limit (reference temperature). The current
through the PTC under normal operating conditions is sufficiently low. At this
current level, the power dissipated by the PTC is low enough, such that the PTC
temperature does not exceed the reference temperature. When a short-circuit
or a high current condition occurs, the current through the PTC increases. The
2.13. CIRCUIT PROTECTION DEVICES 59
power dissipation in the PTC causes an increase in the temperature beyond the
reference temperature, and PTC trips to high resistance state, hence limiting
the current flowing in the circuit. To speed up the temperature rise, such PTC’s
are usually covered by a heat insulator. When the short-circuit or high-current
condition is removed, PTC cools down and returns to its normal low-impedance
state.
PTC’s are usually specified by two current parameters. Rated current (IN )
is the current level, below which the PTC reliably remains in low resistance
mode. Switching current (IS ) is the level beyond which the PTC reliably trips
to high resistance mode. Another parameter of significance is RN , the resistance
of PTC at low resistance mode. PTC thermistors are connected in series to the
circuit to be protected.
NTC thermistors
NTC thermistors can be used as an inrush current limiter device in power supply
circuits. Since the power supply capacitor is initially discharged, a very large
current can flow in the rectifier diodes and this large transient current may
destroy the diodes. If NTC thermistors are connected in series with the diodes,
they present a high resistance at the initial turn-on, while they are cold. This
prevents the large inrush current, saving the diodes. As currents flow through
them, they heat up and become much lower resistance. Therefore, they do not
dissipate a significant power in the steady-state. It is common to put a heat
preserving cover around them to keep them hot and hence low resistance.
NTC thermistors are also used as temperature sensors in many applications.
The measured resistance can be converted to temperature if the resistance versus
temperature characteristics is known.
PTC1 Fuse
Vline Vout Vline Vout
VR1 VR1
(a) (b)
Figure 2.47: Over-voltage protection circuits: Using (a) a PTC and a varactor,
(b) a fuse and a varactor
chosen such that, when there is no over-voltage, the voltage across V R1 is in the
normal range and the current through P T C1 is less than IN . In this case, P T C1
exhibits a low resistance, and V R1 exhibits a very high resistance. When an
over-voltage occurs on the line voltage (for example, due to a flash of lightning in
a thunderstorm), the voltage across V R1 increases beyond its operating range.
The current through V R1 increases rapidly due to the nonlinear nature of the
varistor resistance. This current passes through P T C1 and warms the PTC
up. When this current exceeds IS , P T C1 switches to high impedance mode
isolating the line from the output. To speed up the warming and hence the
2.14. ELECTROMECHANICAL SWITCHES 60
response time, PTC’s are usually placed in a thermally insulating jacket. When
the overvoltage condition is over, PTC cools down and returns to its original
low resistance state. This kind of protection circuits is always present in the
line-voltage inputs of most modern power supplies and power adapters.
Note that a PTC placed in series with a circuit can also be used as an over-
current protection circuit. If the current in the circuit exceeds the predetermined
level, for example, due to an accidental short-circuit, PTC heats up increasing
its resistance, limiting the current.
The circuit shown in Fig. 2.47(b) has a fuse instead of a PTC. A fuse is a
metal wire placed in a glass tube with metal caps on both ends. Fuse metal
melts when too much current flows through it, thereby interrupting the current.
Glass allows a visual inspection of the fuse. Unlike PTC, when a fuse blows, it
does not recover. It has to be replaced with a new fuse.
Their symbols are shown in Fig. 2.48, which also describes their electrical dif-
ferences. In a double-pole or a multi-pole switch, two or more switches operate
in tandem.
NC
NO
NC
NO
Figure 2.49: Symbol of a relay with a DPDT switch with normally-open (NO)
and normally-closed (NC) contacts.
Figure 2.50: Different relays from left to right: SPST, DPDT, 6PDT.
2.15 Examples
Example 14
Using the current divider formula, find the current I1 in Fig. 2.51 with the
reference direction shown.
I1 1K
18mA 2K
3K
Solution
Using the current divider formula of Eq. 2.35, we find
1+3
I1 = −18 = −12 mA
1+3+2
Note that the negative sign arises due to direction of current source versus the
reference direction of I1 .
Example 15
Using nodal analysis find the voltage V1 in Fig. 2.52. Then determine the current
I1 in the reference direction shown.
2mA
1K I1
+ + 3K
5V V1
2K 6V
- +
Solution
We define the bottom wire as ground and write the node equation at V1 as ({V,
mA, kΩ} is a consistent unit set)
V1 − 5 V1 V1 − (−6)
+ + −2=0
1 2 3
2.15. EXAMPLES 63
Solving V1
1 1 6
V1 1 + + = 5 − + 2 = 5 ⇒ V1 = 2.72 V
2 3 3
Example 16
In Fig. 2.53, find the value of the voltage source, V , such that I1 =0.
3K V1 2K V2
I1 1K 3mA
V 4K
+
Solution
We define the bottom wire as ground and write the node equations at V1 and
V2
V1 − (−V ) V1 V1 − V2
+ + =0
3 4 2
V2 − V1 V2
+ −3=0
2 1
Rearranging the equations
1 1 1 1 V
V1 + + − V2 =−
3 4 2 2 3
1 1
−V1 + V2 +1 =3
2 2
Multiplying the first equation by 3 and adding to the second one, we eliminate
V2 to get
39 1
V1 − = −V + 3
12 2
To make I1 , we must have V1 =0, hence
−V + 3 = 0 ⇒ V = 3 V
2.15. EXAMPLES 64
5mA 5K
5K 2K
V1 V2
+
1K 10v 1K 2K
Example 17
Find V1 − V2 in the circuit given in Fig. 2.54.
Solution
Apply nodal analysis method to find V1 and V2 . KCL at V1 node gives
V1 V1 − 10 35
+ − 5 mA = 0 or V1 = = 5.8 V
1K 5K 6
KCL at V2 node gives
V2 V2 − 10
+ + 5 mA = 0 or V2 = 0
2K 2K
Therefore, V1 − V2 = 5.8 V.
Example 18
Assuming that vC (0)=2 V, find i1 (t) in Fig. 2.55(a).
1K v1 2K v2 1K v1 2K v2
+ i1 + i1 +
+ 4µF 5K 5K
5V 4V vC 3mA 5V 4V 2V 3mA
+ +
(a) vC(0)=2V (b)
1K v1 2K v2 1K v1 2K v2
+ i1
5V 4V 5K 3mA Req 5K
+
(c) (d)
Solution
For t=0 solution, we replace the capacitor with a voltage source of 2 V as in
Fig. 2.55(b), and find i1 (0)
v1 − v2 (−4) − 2
i1 (0) = = = −3 mA
2 2
For t=∞ solution, we replace the capacitor with an open circuit as shown in
Fig. 2.55(c) and apply nodal analysis to find v2 (∞)
v2 − v1 v2 v2 − (−4) v2 10
+ −3= + − 3 = 0 ⇒ v2 (∞) =
2 5 2 5 7
Therefore, the current i1 (∞) is found as
v1 − v2 −4 − 10/7 19
i1 (∞) = = =−
2 2 7
We find the equivalent resistance, Req , across the capacitor by killing all sources
as depicted in Fig. 2.55(d)
5×2 10
Req = 5 ∥ 2 = = KΩ
5+2 7
Hence the time constant, τ , is τ =(10/7)KΩ × (4 µF)=(40/7) ms. Now, we can
write the solution for i1 (t) as
19 19 −t/(40/7)
i1 (t) = i1 (∞) + (i1 (0) − i1 (∞))e−t/τ = − + −3 + e mA
7 7
where t is in ms. Note that 5 V voltage source and its 1 K series resistance has
no effect on the value of i1 (t).
Example 19
Assuming that iL (0)=20 mA, find v1 (t) in Fig. 2.56(a).
Solution
For t=0 solution, we replace the inductor with a current source of 20 mA as in
Fig. 2.56(b), and write the nodal equation at vA as
vA − 10 vA
+ 20 + + 25 = 0
0.15 0.1
We find vA =1.3 V. Therefore, v1 (0)=10−1.3=8.7 V. For t=∞ solution, we re-
place the inductor with a short circuit as shown in Fig. 2.56(c) and find v2 (∞)
easily as
v1 (∞) = 10 − 0 = 10 V
We find the equivalent resistance, Req , across the inductor by killing all sources
as depicted in Fig. 2.56(d)
+ + v - + + + v - +
1 1
10V iL 100 2V 10V 100 2V
20m
5mH 200 200
(a) (b)
(c) (d)
Hence the time constant, τ , is τ =5 mH/0.056 KΩ=89.3 µs. Now, we can write
the solution for v1 (t) as
where t is in µs.
Example 20
In Fig. 2.57(a), find the capacitor voltage, vC (t), and current, iC (t), for t > 0.
We have vC (0)=7 V and the switch S is closed at 25 ms.
2K S 2K S
+ iC + +
+
10V vC 10µF 6K 5mA 10V 7V 6K 5mA
iC
(a) (b)
2K S 2K S
+ +
10V vC 6K 5mA Req 6K 5mA
(c) (d)
2K S 2K S
+ + + +
10V 9.14V 6K 5mA 10V vC 6K 5mA
iC
(e) (f)
Solution
We need the find the capacitor voltage vC (25 ms) to act as the initial condition
for the circuit when S is closed. We have switch S open for 0 < t < 25 ms.
For t=0 solution, we replace the capacitor with a voltage source of 7 V as in
Fig. 2.57(b), and find
10 − 7
iC (0) = = 1.5 mA
2
For t=∞ solution (we assume that the switch S is still open!), we replace the
capacitor with an open circuit as shown in Fig. 2.57(c) and find vC (∞) easily
as
vC (∞) = 10 V
We find the equivalent resistance, Req1 , across the capacitor by killing all sources
as depicted in Fig. 2.57(d)
Req1 = 2 KΩ
Hence the time constant, τ1 , for this time duration is τ1 =(2 KΩ)(10 µF)=20 ms.
Now, we can write the solution for vC (t) and iC (t) during 0 < t < 25 ms as
vC (t) = vC (∞) + (vC (0) − vC (∞))e−t/τ1 = 10 + (7 − 10)e−t/20 = 10 − 3e−t/20 V
iC (t) = iC (∞) + (iC (0) − iC (∞))e−t/τ1 = 0 + (1.5 − 0)e−t/20 = 1.5e−t/20 mA
where 0 < t < 25 is in ms. Now, we can find the initial condition for the case S
is closed.
vC (25) = 10 − 3e−25/20 = 9.14 V
The capacitor voltage will be preserved after the switch is closed. This is not
the case for the capacitor current. We find the capacitor current just before the
switch S is closed, t=25−, as
iC (25−) = 1.5e−25/20 = 0.43 mA
At the moment S is closed, we have the circuit as shown in Fig. 2.57(e). We
find the capacitor current at this moment, t=25+, from Fig. 2.57(e) by writing
the node equation as
9.14 − 10 9.14
+ iC + + 5 = 0 ⇒ iC (25+) = −6.09 mA
2 6
For t=∞ solution while S is closed, we open circuit the capacitor as depicted in
Fig. 2.57(f), and write the node equation to find vC (∞) as
vC − 10 vC
+ + 5 = 0 ⇒ vC (∞) = 0
2 6
The equivalent resistance, Req2 and the time constant, τ2 , for t > 25 ms is found
as
Req2 = 2 ∥ 6 = 1.5 KΩ and τ2 = (10 µ)(1.5 K) = 15 ms
Therefore, we can write the capacitor voltage and current as
vC (t) = 0 + (9.14 − 0)e−(t−25)/τ2 = 9.14e−(t−25)/15
iC (t) = 0 + (−6.09 − 0)e−(t−25)/τ2 = −6.09e−(t−25)/15
for 25 < t < ∞ in ms. The capacitor voltage and current is plotted in Fig. 2.58.
Note that the capacitor voltage is always a continuous function, while the ca-
pacitor current may be discontinuous.
2.15. EXAMPLES 68
10
vC (t)
8
i (t)
C
-2
-4
-6
-8
0 5 10 15 20 25 30 35 40 45 50
t (ms)
Example 21
Suppose that the switch in Fig. 2.59 is open for t <0, and closed at t = 0. Find
iL and v1 as a function of time for −∞ < t < ∞.
t=0
v1 200
iL +
3mA 1mH 10v
Solution
The circuit contains only one inductor; it is a first-order RL network. Hence we
can use the method given in Section 2.11.
1. For t > 0 the switch is closed. Killing the sources: The voltage source is
short-circuited, the current is open-circuited. In this case, the resistance
seen by the inductance is 200 Ω.
2. The time constant of the network is τ = L/R = 10−3 /(200)=5 µs.
3. For t < 0, iL (t) = −3 mA, v1 (t) = 0V. At t = 0, the switch is closed, and
iL should be continuous: iL (0+ ) = iL (0− ) = −3 mA. At t = 0+ , KCL
at v1 node implies that there is no current in 200 Ω resistor. Therefore,
v1 (0+ )=10 V.
10
4. Short-circuit the inductor. We find v1 (∞) = 0 and iL (∞) = 200 −3 =
47 mA
5. We write the solutions for t > 0 as
Example 22
Considering the circuit in Fig. 2.60, at t = 0, the switch is opened, and 1 kΩ
resistor is in the circuit. Find the voltage vo (t).
50mH 10
L iL(t)
+ vo(t)
12V t=0
1K
Solution
1. 1 kΩ resistor is shorted for t < 0, while 10 Ω resistor is in the circuit. We
have vo (t) = 0 for t < 0. The current iL (0− ) just before the switching,
can be found by short-circuiting the inductor: iL (0− ) = 12/10 = 1.2 A.
2. Since the inductor current is continuous, the inductor current just after
the switching, iL (0+ ) = 1.2 A. Hence vo (0+ ) = 1.2 · 1000 = 1200 V!
3. At t = ∞, we short-circuit the inductor. We determine
iL (∞) = 12/1010 = 11.9 mA and vo (∞) = 11.9 V.
1200
1000
800
v (t) (V)
600
o
400
200
0
0 50 100 150 200 250 300 350 400
t (µs)
2.16 Problems
(Answers of most problems are in p. 311.)
1. Find the values and tolerances of resistors with the following color codes:
(a) green-blue-red-silver
(b) orange-white-yellow-silver
(c) violet-green-brown-gold
2. Find the average power delivered or dissipated by the sources and resistor
in Fig. 2.62. Make sure that the average power delivered is equal to the
average power dissipated.
IV
+V
R IR
5 cos ωt 33Ω 3
3. Find the rms value of the periodic voltage waveform shown in Fig. 2.63.
5
v(t)
-5
82
470 22K 3.9K
47
Req 51 51 Req 12 Req 12K 15K 6.8K
470 470 1K
Req 1K 2.2 K Req 51 2.7 K 330
(d) (e)
1K 2K
+ +
12V 3K V2
-
+
4K 5K V1
-
8. Using the voltage divider formula, find the voltages V1 and V2 in Fig. 2.65.
9. Determine the current I1 in Fig. 2.66. Note that {V, µA, MΩ} is consistent
unit set.
1.2M 820K
I1 +
1.5M 2µA 2.2M 3V 560K
10. Find the marked variables using nodal analysis of the circuits in Fig. 2.67
(three significant figures). Check your results using LTSpice.
11. In Fig. 2.68, find and plot iL (t) and vR (t) with iL (0)=30 mA. Initially S
is closed, at t1 =40 µs, S is opened. Hint: You need to find the inductor
current, iL (t1 ) to act as the initial condition for the circuit when S is
opened.
2.16. PROBLEMS 73
S 15K
+v - +
R
90mA 20mH iL 12V
1K
12. Find and plot iL (t) in Fig. 2.69 with iL (0)=5 mA.
13. Consider the circuits in Fig. 2.70(a) and (b). Both circuits are driven by
a step current source iS (t), shown in Fig. 2.70(c). Find and sketch iC (t),
vC (t), iL (t), and vL (t). Assume that vC (0) = 0 and iL (0) = 0. Verify
your results with LTSpice.
14. Find the resistance of 500 m copper wire of diameter 0.1 mm.
100 5mH
vIN + 20mH iL
− 1K 150
vIN
10
100µs
t
-10
21. Considering the circuit given in Fig. 2.74, find and plot the current in the
inductor, iL (t), as a function of time between 0 < t < 10 µs.
2.16. PROBLEMS 75
-15 V
R1 C1
Vout + vout(t)
R2 12 sin(ωt) C2
(a) (b)
Vin
1K iL 3V
+
+ 2K vL
Vin − t
L=3mH -2V
22. For the circuit of Fig. 2.75, find and plot the voltage of the capacitor,
vC (t), as a function of time between 0 < t < 30 ms.
2.16. PROBLEMS 76
vIN iL(t)
+
10V vIN 100µH
iL(0)=0
0 5µs t
iin(t)
5 mA 10µF +
iin(t) vC(t)
vC(0)=-2
0 12 ms t
AUDIO CIRCUITS
The most natural way of communication for people is to speak to each other.
The voice is transmitted and received in electronic communications, to enable
people communicate over large distances. The first thing that must be done
is to convert voice into an electrical signal, and process it before transmission.
The last process in a transceiver, on the other hand, is to recover voice from the
received RF signal. The audio circuits of TRC-11 are discussed in this chapter.
The mathematical tools necessary to analyze circuits used in TRC-11 are also
developed.
English electrical engineer Oliver Heaviside (1850–1925) adapted complex
numbers to analyze electrical circuits. Complex numbers are very important in
Electrical Engineering, so we give a brief summary of complex numbers here.
and a and b are real numbers. j is called the imaginary unit and has the property
of j 2 = −1.* a is the real part and b is the imaginary part:
This complex number can be shown as a vector in the complex plane as demon-
strated in Fig. 3.1(a). Two complex numbers a + jb and c + jd are equal if and
only if a = c and b = d. Real numbers are a subset of complex numbers. If the
real part of a complex number is zero, then it is called an imaginary number.
For example, 2 + j0 and −11 + j0 are real numbers, while 0 − j6 is an imaginary
number.
The complex conjugate of a complex number z = a + jb is z ∗ = a − jb.
* Electrical Engineers prefer to use the symbol j rather than i, since the symbol i is reserved
for current.
Figure 3.1: (a)Vector showing the complex number z = a + jb, (b) addition of
two complex numbers, (c) subtraction of two complex numbers.
The algebra of complex numbers is the same as the algebra of real numbers with
j 2 replaced by −1:
Addition (see Fig. 3.1(b)):
Multiplication:
Division:
z1 a + jb ac + bd bc − ad
= = 2 +j 2
z2 c + jd c + d2 c + d2
1 1 c d
= = 2 2
−j 2
z2 c + jd c +d c + d2
Absolute value: p
|a + jb| = a2 + b2
If z = a + jb, then
zz ∗ = |z|2 = a2 + b2
We also have
|z1 z2 | = |z1 ||z2 |
and
z1 |z1 |
= if z2 ̸= 0
z2 |z2 |
There is a relation between sinusoids and exponential function as follows:
This is called Euler’s formula. In other words, cos ϕ is the real part of ejϕ , and
sin ϕ is the imaginary part. Sinusoids can be expressed as
cos ϕ = Re ejϕ
(3.4)
3.2. PHASORS 79
or
ejϕ + e−jϕ
cos ϕ = (3.5)
2
and
sin ϕ = Im ejϕ
(3.6)
or
ejϕ − e−jϕ
sin ϕ = (3.7)
2j
in turn. The magnitude of this exponential function is
|ejϕ | = 1 (3.8)
z = rejϕ (3.10)
z1 z2 = r1 r2 ej(ϕ1 +ϕ2 )
or
z1 z2 = r1 r2 ∠(ϕ1 + ϕ2 )
the division as
z1 r1
= ej(ϕ1 −ϕ2 )
z2 r2
or
z1 r1
= ∠(ϕ1 − ϕ2 )
z2 r2
Clearly, the multiplication and division are easily performed in exponential or
polar forms, while the addition and subtraction are easier in rectangular form.
3.2 Phasors
In Electrical Engineering we frequently deal with sinusoidal signals; sometimes
we add them, sometimes we subtract them from each other. Consider two
sinusoids at the same frequency ω, but with a differing amplitude and phase.
The sum of these sinusoids can be written as
A cos(ωt + θ1 ) + B cos(ωt + θ2 ) =
p
A sin θ1 +B sin θ2
= A2 + B 2 − 2AB cos(θ2 − θ1 ) cos ωt + tan−1 A cos θ1 +B cos θ2 (3.13)
3.2. PHASORS 80
Just the sum of two sine waves at the same frequency is a rather clumsy equation.
It is obviously difficult to manipulate equations of sine waves using the time-
domain notation above.
We can simplify a great deal if we use the Euler’s formula. From Eq. 3.4 we
can write
A cos(ωt + θ1 ) = Re{Aej(ωt+θ1 ) } = Re{Aejωt ejθ1 } (3.14)
similarly
B cos(ωt + θ2 ) = Re{Bej(ωt+θ2 ) } = Re{Bejωt ejθ2 } (3.15)
jωt
To simplify the notation, we can get rid of the e term since the frequency is
common. To simplify further we can also get rid of the real part operator Re:
Example 1
Below are some examples of conversion from time-domain to phasor notation.
o
5 cos(ωt + 23o ) ⇒ 5ej23
o o
10 cos(ωt + 12o ) + 8 cos(ωt − 76o ) ⇒ 10ej12 + 8e−j76 =
= 10(cos 12o + j sin 12o ) + 8(cos (−76o ) + j sin(−76o )) =
= 9.78 + j2.08 + 1.93 − j7.76 = 11.71 − j5.68
in the last example, we used Eq. 3.9 to convert from the polar form to rectangular
form.
In the following examples, we use Eq. 3.18 to convert the phasors to time
domain: o
7e−j12 ⇒ 7 cos(ωt − 12o )
√ √
6 + j6 = 6 2ejπ/4 ⇒ 6 2 cos(ωt + π4)
−1 o
3 − j4 = 5ej tan (−4/3)
= 5e−j53.1 ⇒ 5 cos(ωt − 53.1o )
In the last example, we first used the formulas of Eq. 3.12 to convert the rect-
angular form complex number to the exponential form.
Imag Imag
IC=VCjωC VL =ILjωL
VC Real IL Real
(a) (b)
Figure 3.2: Current and voltage phasors for a capacitor (a) and for an inductor
(b).
Note that the resulting phasor equation is like Ohm’s law with 1/(jωC) replac-
ing R. The current and voltage phasors for a capacitor are demonstrated in
Fig. 3.2(a).
We note that real-life capacitors behave like that given in Eq. 3.25 only
up to a frequency limit (SRF) due to the inductance of the capacitor leads
or internal connections. Above this frequency limit, the capacitor acts like a
small inductor! Since small capacitors have higher SRF, it is recommended to
choose the smallest capacitor that will satisfy the requirements. To extend the
frequency range of capacitors, often small capacitors with higher SRF is placed
in parallel with larger capacitors.
If input signals x1 (t) and x2 (t) (voltage or current) yield the output signals
y1 (t) and y2 (t), respectively, then a linear combination of inputs, ax1 (t) + bx2 (t)
yields the same combination of the individual outputs, ay1 (t) + by2 (t), where a
and b are real numbers.
SRF is the self-resonance frequency (SRF). It can be found in the data sheets of capacitors.
The self-resonance frequency of inductors can be found in the data sheets of off-the-shelf
inductors or it can be measured for an in-house manufactured inductor.
3.3. LINEAR CIRCUITS 83
Figure 3.3: A linear circuit block with an input signal x(t) and an output signal
y(t).
If we apply a combination of two inputs ai1 (t) + bi2 (t), then the total voltage
developed across the resistor is
v(t) = R[ai1 (t) + bi2 (t)] = aRi1 (t) + bRi2 (t) = av1 (t) + bv2 (t) (3.28)
v1 = 0 and v2 = 0 (3.31)
We are not limited to first-order circuits with just one capacitor or one inductor.
We can deal with any number of capacitors and inductors.
The phasor method is not able to find the transient solution that occurs right
after the sinusoidal signal is applied. Rather, it can find the steady-state solution
long after the sinusoidal signal is applied and transients have disappeared. Note
that this method is applicable only when the excitation to the circuit sinusoidal.
Moreover, the circuit should not contain any nonlinear elements like diodes.
Example 2
Consider the circuit shown in Fig. 3.4(a) and find the voltage across the capac-
itor.
Since the excitation is sinusoidal and the circuit contains only linear elements
we can use the procedure above:
1K 1K
+ + + +
5cos100t 5µF vC 5 -j2K VC
- -
(a) (b)
Hence
−j2K −j2(1 + j2) √ o
VC = 5= 5 = 4 − j2 = 20e−j26.5
1K − j2K 1+4
√
4. In time domain, we have vC (t) = 20 cos(100t − 26.5o ) V
Example 3
Consider the circuit shown in Fig. 3.5(a) and find the voltage v1 (t).
v1 2µH V1 j20
3sin(107t) 50 50
mA 3nF -j3
mA -j33.3
(a) (b)
We have a linear circuit excited with a sinusoidal signal. We can use the
phasor method:
1. Replace the inductor with jωL = j107 × 2 × 10−6 = j20, and the capacitor
with 1/(jωC) = 1/(j107 × 3 × 10−9 ) = −j33.3
2. Replace the current source with the phasor −j3 × 10−3 .
3. Referring to Fig. 3.4(b), the voltage phasor across the capacitor can be
found from nodal analysis:
V1 V1 50 − j13.3
−j3 × 10−3 = + = V1
50 j20 − j33.3 −j13.3 × 50
Hence, we find
−1.99 1.99∠180o
V1 = = = 0.038∠194.9o
50 − j13.3 51.7∠ − 14.9o
where we converted the nominator and the denominator from rectangular
form to polar form to simplify complex division.
4. In time domain, we have v1 (t) = 0.038 cos(107 t + 194.9o ) V
In general, for circuits where the current and voltage may have phase
difference, the power dissipated in the circuit can be written as
∗ ∗
VI V I
P = Re = Re (3.36)
2 2
3.3.3 Impedance
We have observed that the definition of phasors allowed us to convert the dif-
ferential relations in time into algebraic relations in angular frequency.
3.3.4 Admittance
3.3. LINEAR CIRCUITS 87
The unit of Y is Siemens (S), named after German inventor Ernst Werner von
Siemens (1816–1892). The unit of the admittance is also commonly referred to
as mho, and its symbol 0 (an upside-down Ω). The real part of the admittance,
G, is conductance and the imaginary part, shown by B, is the susceptance.
In circuits with many parallel components, it is more convenient to use
admittance rather than impedance: Admittances of parallel circuits can simply
be added.
Example 4
Let us find the impedance of the series connected circuit of Fig. 3.6(a) and the
admittance of the parallel connected circuit of Fig. 3.6(b) at the given frequen-
cies.
10 820nH f=5 MHz
R1 L1 C1 L2 C2
Z1 = ? Y2 = ? R2
f=20 MHz 47pF 470 10µH 82pF
(a) (b)
1
Z1 = R1 + jωL1 + = 10 + j(1.26 · 108 )(820 · 10−9 )
jωC1
1
+ = 10 + j103 − j169 = 10 − j66 Ω
j(1.26 · 10 )(47 · 10−12 )
8
1 1 1
Y2 = + jωC2 + = + j(3.14 · 107 )(82 · 10−12 )
R2 jωL2 470
1
+ == 2.13 · 10−3 + j2.58 · 10−3 − j3.18 · 10−3 =
j(3.14 · 10 )(10 · 10−6 )
7
= 2.13 − 0.61 m0
Example 5
Consider the circuit given in Fig. 3.7. The impedance can be found by applying
a voltage phasor V and finding I in terms of it:
V V
I= +
100 jω(2 · 10−6 ) + 1/(jω(3 · 10−9 ))
3.3. LINEAR CIRCUITS 88
I 2µH
+
V 100 3nF
-
Note that the real part of the impedance acts like a frequency dependent resistor.
On the other hand, the admittance is found as
Vo (ω)
H(ω) = (3.40)
Vi (ω)
Example 6
To find the transfer function of the first-order circuit shown in Fig. 3.8(a), we
use the procedure of page 89 and write
10µΗ 100 dB
-3.52
Vin 200 Vo -20dB/dec
4.77MHz
(a) (b)
Figure 3.8: (a) A first-order circuit for transfer function evaluation, (b) Transfer
function.
Example 7
To find the transfer function of the first-order circuit shown in Fig. 3.9(a), we
use the same procedure:
1. Using the voltage divider formula, we find
Vo 1K + 1/(jω(2.2n)) 1 + jω (2.2n)(1K)
H(ω) = = =
Vin 22K + 1K + 1/(jω(2.2n))) 1 + jω (2.2n)(23K)
(3.46)
The transfer function fits the last generic form of Eq. 3.42.
3.4. TRANSFER FUNCTION 91
dB
22K 0 -20dB/dec
1K
Vin Vo 72 kHz
3.1 kHz
2.2n -27
(a) (b)
Figure 3.9: (a) A first-order circuit for transfer function evaluation, (b) Transfer
function.
we can draw the mid-frequency asymptote with slope −20 dB/dec accord-
ingly.
5. The transfer function is drawn approximately as in Fig. 3.9(b) using the
three asymptotes.
Example 8
Consider the network given in Fig. 3.10. Let us find the transfer function and
plot the magnitude of the transfer function. We apply a voltage phasor at the
input and find the output phasor using nodal analysis:
1/(jω(3 · 10−9 )) 1
H(ω) = −6 −9
=
10 + jω(2 · 10 ) + 1/(jω(3 · 10 )) 1 − ω (6 · 10 ) + jω(3 · 10−9 )
2 −15
3.5. THÉVENIN EQUIVALENT CIRCUIT 92
+ 10 2µH +
Vi 3nF Vo
- -
−5
−10
|H(ω)|dB
−15
−20
−25
−30
−35
−40
5 6 7
10 10 10
f (Hz)
can be modelled with just two components at the same terminals: A voltage
source, vth and a series resistance, Req as shown in Fig. 3.12(b). This simple
model is called the Thévenin equivalent circuit, named after the French engi-
neer Léon Charles Thévenin (1857–1926). Since the circuit is linear, its V − I
Req
A A
+ Vth +
− −
B B
(a) (b)
1. Find the voltage between the terminals A and B while those terminals
are open-circuited to determine vth .
2. Kill the voltage and current sources within the black-box: Short-
circuit the voltage sources and open-circuit the current sources. Find
3.5. THÉVENIN EQUIVALENT CIRCUIT 94
Example 9
Consider the circuit given in Fig. 3.13(a). Let us find the Thévenin equivalent
of the circuit inside the dashed box. Applying the procedure:
A
+ A
+ 39 + 39 vth
− 33 56 − 33
20V 1A B 20V 1A - B
(a) (b)
A Req A
39 Req vth + 17.8
33 − 56
B 27V B
(c) (d)
1. Find the voltage phasor between the terminals A and B while nothing
is connected to those terminals to determine the phasor Vth (ω).
3.6. NORTON EQUIVALENT CIRCUIT 95
2. Kill the voltage and current sources within the black-box: Short-
circuit the voltage sources and open-circuit the current sources. Find
the impedance between the terminals A and B to determine Zeq (ω).
Example 10
Consider the RLC circuit given in Fig. 3.14(a). Since the excitation is sinusoidal,
we can find the Thévenin equivalent circuit: Applying the procedure:
A A
+ + +
R R
C L RL A C L Vth
Acos(ωt) B B
(a) (b)
A A
+
R Zeq + Zeq
C L Vth RL
A B B
(c) (d)
V V V −A jωL
+ + =0 or Vth (ω) = A
jωL 1/(jωC) R R− ω 2 RLC + jωL
2. Kill the voltage source phasor as in Fig. 3.14(c). Find the impedance
between the terminals A and B (also with RL removed):
jωRL
Zeq (ω) =
R− ω 2 RLC + jωL
Clearly, for the same black-box Req is the same value for Thévenin and
Norton equivalent circuits, because they are found in the same way. Since they
represent the same straight line we also have vth = Req iN .
Example 11
Consider the same circuit example of Thévenin equivalent circuit in Fig. 3.15(a).
Norton equivalent circuit is found by:
A A
+ 39 + 39 iN
− 33 56 − 33
20V 1A B 20V 1A B
(a) (b)
A A
39 iN
33 Req 17.8 Req 56
B 1.51A B
(c) (d)
1. Using nodal analysis, we find the short-circuit current between the termi-
nals A-B (Fig. 3.15(b)):
20
iN = 1 + = 1.51A
39
2. Kill the sources as in Fig. 3.15(c). Find the resistance between the termi-
nals A and B (with 56 Ω removed): Req = 33 ∥ 39 = 33 · 39/(33 + 39) =
17.8.
The Norton equivalent circuit is shown in Fig. 3.15(d). For this example, finding
the Norton equivalent circuit was simpler than finding the Thévenin equivalent.
If there are parallel elements across the terminals A and B, Norton equivalent
circuit should be preferred, since it gets rid of those elements when the terminals
3.6. NORTON EQUIVALENT CIRCUIT 97
are shorted. On the other hand, if there are series elements at the terminals
A or B, Thévenin should be preferred. The series elements will be eliminated
when the terminals are open-circuited.
2. Kill the voltage and current sources within the black-box: Short-
circuit the voltage sources and open-circuit the current sources. Find
the impedance between the terminals A and B to determine Zeq (ω).
Example 12
Refer to the RLC circuit given in Fig. 3.16(a) considered earlier. We can find
the Norton equivalent circuit as
A A
+ +
R R IN
C L RL A C L
Acos(ωt) B B
(a) (b)
A A
+
R Zeq
C L IN Zeq RL
A B B
(c) (d)
Figure 3.17: (a) Two circuit pieces connected to each other by means of two
terminals, (b) Thévenin equivalent of Circuit 1.
obtained, Circuit 1 can be replaced by its equivalent as in Fig. 3.18 and Circuit
2 can be analyzed.
Example 13
Consider the circuit in Fig. 3.19. Voltage across 2.7 kΩ resistor can be found in
many different ways.
3.6. NORTON EQUIVALENT CIRCUIT 99
Figure 3.19: Example showing the use of Thévenin and Norton equivalent cir-
cuits.
and short it. The short circuit current is the equivalent current and the equiva-
lent resistor is 0.72K+2.2K. In third step, we notice that the equivalent current
and 1 mA current source are in parallel. Combining them in a single source
of 3.47 mA, we convert the Norton circuit to its Thévenin equivalent. After
another Thévenin equivalent conversion, output voltage is obtained as 0.75 V.
Step 1:
1.2K 2.2K 3.3K 1.5K 1.2K 0.72K
+ + +
12V 1.8K 1K 2.7K 12V 1.8K 7.2V
1mA
Step 2:
0.72K 2.2K 3.3K 1.5K 0.72K 2.2K
+ +
7.2V 1K 2.7K 7.2V 2.92K
1mA 2.47mA
Step 3:
3.3K 1.5K 2.92K
+
2.92K 1K 2.7K 2.92K 10.1V
2.47mA 1mA 2.47mA 1mA
+ +
1.4V 2.7K 0.75V
Example 14
Refer to the resistive circuit depicted in Fig. 3.21(a). Use the superposition
principle to find the voltage vA (t).
R1 R3 vA R1 R3 vA
+ R2 + R2
v1(t) − R4 v1(t) − R4
i2(t)
(a) (b)
vA R1 R3 vA
vth(t) + Req R2
− R4 R4
i2(t)
(c) (d)
Solution
1. Kill the current source as shown in Fig. 3.21(b). Find the Thévenin equiv-
alent circuit of the circuit in the dashed box. We find vth using the voltage
3.7. SUPERPOSITION PRINCIPLE 101
divider formula (Eq. 2.30) and Rth is found by parallel and series resistor
combination formulas:
R2 R1 R2
vth (t) = v1 (t) and Req = + R3
R1 + R2 R1 + R2
vA1 (t) for the voltage source is found from the voltage divider in Fig. 3.21(c):
R4 R4 R2
vA1 (t) = vth (t) = v1 (t)
Req + R4 Req + R4 R1 + R2
2. Kill the voltage source as in Fig. 3.21(d). We find the total resistance
across the current source, RT , and hence the corresponding vA2 (t) as
Req R4
RT = Req ∥ R4 = and vA2 (t) = −RT i2 (t)
Req + R4
R4 R2
vA (t) = vA1 + vA2 = v1 (t) − RT i2 (t)
Req + R4 R1 + R2
As an exercise, solve the same circuit, using nodal analysis. As a third alterna-
tive, you can use Norton equivalent circuit for the part in dashed lines to get
two current sources and two resistors in parallel which are easily combined.
Example 15
Consider the first-order circuit with two sources and with an initial value given in
Fig. 3.22(a). Find the capacitor voltage, vC (t) using the superposition principle.
1K 3K 1K 3K
vC(0)=5V vC(0)=0
+ +
12V 3mA 12V
+ 5µF + 5µF
(a) (b)
1K 3K 1K 3K
vC(0)=0 vC(0)=5V
+ +
3mA
5µF 5µF
(c) (d)
Solution
We use the superposition principle with two sources and one initial condition.
1. Kill the current source and the initial condition as shown in Fig. 3.22(b).
The final value of the capacitor voltage is vf =−12 V. The time constant
is τ =(5µ)(1K+3K)=20 ms. The solution for this case is
2. Kill the voltage source and the initial condition as depicted in Fig. 3.22(c).
The final value of the capacitor voltage is determined by the 3 mA current
flowing in 1K resistor: vf =3 V. The time constant, τ , remains the same.
The corresponding solution is
vc (t) = 3 + (0 − 3)e−t/20m
3. Kill the voltage source and current source as shown in Fig. 3.22(d). The
final value of the capacitor voltage is zero. The time constant is the same.
Hence we have
vc (t) = 0 + (5 − 0)e−t/20m
vc (t) = −9 + 14e−t/20m
Example 16
Consider the circuit in Fig. 3.23(a) driven by sinusoidal sources. The two sources
are at two different frequencies, ω1 = 103 /3 and ω2 = 103 : v1 (t) = 2 cos(103 /3t+
π/6) and i2 (t) = 5 cos(103 − π/4). Find the steady-state value of the resistor
current, i3 (t) using the superposition principle.
ω1 circuit
0.8K 0.8K
+ R1 R2 + V1 R1 R2
v1 C i2 i3 jπ/4
i3
2µ 3K 2e 1/(jω1C) 3K
(a) (b)
ω2 circuit
0.8K
R1 I2 R2
5e-jπ/4 i3
1/(jω2C) mA 3K
(c)
Solution
Since the sources are sinusoidal and we need the steady-state solution, we can
use phasors.
1. Kill the current source and replace the circuit with phasor equivalents at
ω1 = 103 /3 as depicted in Fig. 3.23(b). We find the resistor voltage using
the voltage divider formula. Then the current is found:
R2 ∥ (1/jω1 C) 1
I3 = V1 = 0.485ej0.125 mA
R1 + (R2 ∥ (1/jω1 C) R2
2. Kill the voltage source and replace the circuit with phasor equivalents at
ω2 = 103 as depicted in Fig. 3.23(c). We find the resistor current using
the current divider formula.
R1 ∥ (1/jω2 C)
I3 = I2 = 0.653e−j1.687 mA
R2 + (R1 ∥ (1/jω2 C)
3. We convert the phasor to time domain equivalents and add them to find
the steady-state solution of the resistor current
+ + +
vin Avin iin Rfiin
(a) (b)
+
vin Gfvin iin Biin
(c) (d)
Figure 3.24: (a) Voltage amplifier, (b) transimpedance amplifier, (c) transcon-
ductance amplifier, and (d) current amplifier.
+VC
v2 − v2 vo
vo -
+
− A(v1-v2)
v1 + v1 +
-VC
(a) (b)
Figure 3.25: (a) OPAMP symbol, (b) an ideal OPAMP equivalent circuit
at its output. Typically, Vmax is slightly smaller than the positive supply voltage
+VC , and Vmin is slightly greater then the negative supply voltage −VC .
We can express the characteristics of a real OPAMP as follows:
Vmin Vmax
A(v1 − v2 ) if A ≤ vin1 − vin2 ≤
A
vo = Vmax if v1 − v2 > Vmax
A
(3.48)
if v1 − v2 < Vmin
Vmin
A
Typically, A is a large number in the order of 105 or 106 . This implies that
as long as Vmin /A ≤ v1 −v2 ≤ Vmax /A is satisfied v1 −v2 is a very small number.
Hence we can write
v1 ≈ v2 if Vmin ≤ vo ≤ Vmax (3.49)
This approximation simplifies the solution of most OPAMP circuits. In the
solution of OPAMP circuits, the output node vo should be considered as a
voltage source, which may provide the needed current. It is fair to assume that
there is no current going into the OPAMP at the pins v1 and v2 .
R2 R2
vin R1 v2 − R1 v2 −
vo vo
Rin R3 R3
+ vin +
v1 v1
Rin
(a) (b)
v2 vo − v2 vo R2
= or A= ≈1+ (3.51)
R1 R2 vin R1
The input impedance, Rin , of this amplifier is very high (infinity for an ideal
OPAMP), since no current flows through R3 . Similar to the inverting amplifier,
the value of R1 should be chosen in the range 1 kΩ to 100 kΩ.
If we choose R2 = 0 and/or R1 = ∞, then the gain becomes one. Such a
circuit is called unity gain amplifier or voltage follower, and it is commonly used
as a buffer. Although it does not provide any voltage gain to the input signal,
it is used to transfer the input voltage intact to the output while altering the
impedance that appears at the terminals of vin to the low output impedance of
OPAMP. This can provide a large power gain, because the voltage at the source
can now be applied to a relatively low impedance load.
Rin2 R4
R2 R3 Rin2
vin2
R3 v2 −
R1 vin2
vin1 v2 − vo
vo R1
vin1 +
Rin1 v1
+ R2
R4 v1 Rin1
(a) (b)
v2 − vin1 v2 − vin2 v2 − vo
+ + =0
R1 R2 R3
3.9. OPERATIONAL AMPLIFIERS 107
Since v2 ≈ v1 = 0, we find
R3 R3
vo = − vin1 + vin2 (3.52)
R1 R2
If R1 = R2 = R3 , we get the sum of input signals with inverted polarity:
vo = −(vin1 + vin2 ).
The input impedances are given by Rin1 = R1 and Rin2 = R2 , since v2 =0.
Exercise
Design an OPAMP circuit to add four signals (with inverted polarity).
Example 17
Design a circuit with the transfer characteristics vo = 5vin + 3.
Solution
Use a difference amplifier and choose vin1 = vin , R1 = R2 , R4 /R3 = 9 and
vin2 = −1/3V.
Rf
iin v2 −
vo
v1
+
R1 v R2
2 −
vo
v1 +
R3
R4
+
VR IL RL
Using the voltage divider relation, we can find the voltage v2 in terms of vo :
R1
v2 = vo (3.55)
R1 + R2
Using the node equation at node v1 ;
v1 − V R v1 − vo
+ + IL = 0 (3.56)
R4 R3
If vo is not saturated, we have v1 = v2 . Combining these equations, we find
R2 1 1 1 VR
+ − vo − + IL = 0 (3.57)
R1 + R2 R3 R4 R3 R4
With the condition R4 = R1 R3 /R2 , the term inside the square brackets van-
ishes, and we find
VR
IL = (3.58)
R4
IL is independent of RL , so the resulting circuit acts like a current source.
3.9. OPERATIONAL AMPLIFIERS 109
3.9.7 Integrator
An OPAMP can be used to build an integrator as depicted in Fig. 3.30(a). We
observe that v1 = 0, since there is no current in R2 . The current in R1 is equal
vC
+ R1
R1 C vC
vin v2 − vin + v2 −
vo C vo
+ +
R2 v1 R2 v1
(a) (b)
to the current in C:
vin − v2 dvC d(v2 − vo )
=C =C
R1 dt dt
If the output voltage, vo , is not saturated (Vmin ≤ vo ≤ Vmax ), we can use
v2 ≈ v1 = 0. We integrate both sides of the equation above from 0 to t to get
Z t
1
v0 (t) = vo (0) − vin (τ )dτ (3.59)
R1 C 0
where vo (0) is the output voltage at t = 0.
Example 18
Assume that R1 = 1K, C = 10µF, vo (0) = 3V and vin is a 5 V pulse waveform
as shown in Fig. 3.31(a). We find the output voltage as plotted in Fig. 3.31(b).
vin vo
5V
3V
1V
2 6 t (ms) 2 6 t (ms)
(a) (b)
Figure 3.31: (a) Input signal, vin , of the integrator, (b) Output signal, vo , of
the integrator
3.9.8 Differentiator
A differentiator is built from an OPAMP as drawn in Fig. 3.30(b). If the output
is not saturated, we have v2 = 0 since v1 = 0. The current in C is equal to the
3.9. OPERATIONAL AMPLIFIERS 110
current in R1 :
d(vin − v2 ) dvin v2 − vo vo
C =C = =−
dt dt R1 R1
Hence
dvin
v0 = −R1 C (3.60)
dt
C
v2 −
R2 vo
vin R1 v2 − R2
vo +
vin R1 v3 v1
R3 C2
+ C1
v1
(a) (b)
(capital letters) assuming that the input signals are sinusoidal. We know that
V2 ≈ V1 = 0. Hence, we write
Vin −Vo
= − Vo jωC
R1 R2
Therefore, the transfer function is given by
Vo (ω) R2 1
H(ω) = =− (3.61)
Vin(ω) R1 1 + jωR2 C
√
When ωR2 C = 1, the magnitude of low-pass-filter term becomes 1/ 2. This
frequency,
1 1
ω0 = or f0 = (3.62)
R2 C 2πR2 C
is known as the√ corner frequency or cutoff frequency. In decibels the magnitude
is 20 log10 (1/ 2) = −3 dB.
The magnitude of the transfer function in decibels is
V0 R2
= 20 log10 − 10 log10 (1 + (ω/ω0 )2 ) (3.63)
Vin dB R1
The first term is the dB gain of the inverting amplifier. The second term is the
low-pass-filter term. We plot this transfer function in Fig. 3.33 for R2 = 1K
3.9. OPERATIONAL AMPLIFIERS 111
−5
−10
−15
|H(ω)|dB
−20
−25
−30
−35
−40
1 2 3 4 5
10 10 10 10 10
f (Hz)
We will show in p. 227 that the shape of the response becomes desirable if
Fig. 3.33 also shows the magnitude of the transfer function for this filter for
R1 = R2 = 1K and C1 = C2 = 0.159µF. It has a higher slope (−40 dB/decade)
above the corner frequency.
3.10 Microphone
Microphone is a device that converts sound into an electrical signal. Sometimes,
it is abbreviated as mike. Many different types of microphones are available.
Dynamic microphone: The diaphragm of the microphone is attached to a
small movable induction coil, which is positioned in the magnetic field of
a permanent magnet. When the sound wave moves the diaphragm, the
coil moves in the magnetic field, producing a varying current in the coil
through electromagnetic induction. This type of microphone has a low
electrical impedance.
Crystal or piezoelectric microphone: The diaphragm of the microphone
applies a pressure to a piezoelectric crystal, which creates an electrical
voltage proportional to the applied pressure. This type of of microphone
has a high electrical impedance. They are commonly used in electrical
guitars, directly contacting the vibrating surfaces.
Condenser microphone: The basic condenser microphone is a parallel plate
capacitor with one of the plates made from a very thin diaphragm (typi-
cally metallized mylar film) while the other one is a thick metal plate. A
cross-section of this type of microphone is shown in Fig. 3.34 along with
its equivalent circuit.
Vdc
metalized very large
aluminum polyethylene R
ring diaphragm
spacer
+
perforated Co
area A aluminum V(t)
grid
-
insulating
protective thin
grid polyethylene
ring spacer
do
mechanical structure equivalent circuit
Figure 3.34: Condenser microphone and its equivalent circuit with a bias volt-
age.
3.10. MICROPHONE 113
Q = Co Vdc (3.68)
When there is sound incident on the thin diaphragm, sound pressure forces
the diaphragm to vibrate back and forth. Suppose that the spacing is large
enough so that the electrostatic attraction force between the capacitor
plates is negligible compared to the force due to the sound pressure. With
that assumption, the spacing of the capacitor plates varies linearly as a
function of time: d(t) = do ± ∆(t), where ∆(t) is proportional to the
sound signal amplitude. Hence, the value of the capacitor becomes also a
function of time
ϵo A ϵo A Co do
C(t) = = = (3.69)
d(t) do ± ∆(t) do ± ∆(t)
Since the resistor R has a large value, the total charge Q on the capacitor
remains the same. Therefore, the voltage of the capacitor must be a
function of time:
Q Co Vdc ∆(t)
V (t) = = (do ± ∆(t)) = Vdc 1 ± (3.70)
C(t) Co do do
The sound is thus converted to the electrical signal Vdc ∆(t)/do in a lin-
ear manner. For ordinary sound levels, the variation in the membrane
displacement |∆(t)|max is a small fraction of the gap, do . Note that the
sensitivity of a microphone depends on the size of do : smaller do is, more
sensitive the microphone becomes. To preserve linearity, the electrostatic
force between the plates should be negligible, hence the capacitor plate
separation should not be very small. In most commercial microphones, a
compromise between these two requirements is found at about do = 25 µm.
The expression for V (t) in Eq. 3.70 tells that there is a voltage source
Vdc ∆(t)/do additive to the capacitor (i.e. in series with the capacitor).
Fig. 3.34 also shows the overall equivalent circuit.
The microphones must be connected to amplifiers using well-shielded (some-
times double-shielded) cables to prevent hum. Hum is the unwanted low-
frequency sound caused by power-line frequencies (50 or 60 Hz) in audio
systems. Unshielded cables cause power-line frequencies be capacitively
coupled to the sensitive input an amplifier causing an unpleasant noise.
For a microphone with a surface of A = 1 cm2 , and a gap of do = 25 µm,
Co is approximately 35 pF. The capacitance of a typical shielded cable
3.10. MICROPHONE 114
charged
polymer
diaphragm
+
Co
V(t)
-
10mm
front view cross section equivalent circuit
Vdc Vdc
R + R
Co Vout
Vout V(t)
I(t)=gmV(t)
-
electret FET buffer
microphone amplifier equivalent circuit
Figure 3.36: Buffered electret condenser microphone and its equivalent circuit.
The microphone housing contains the microphone and a field effect tran-
sistor (FET), which is used as a buffer amplifier. The internal wiring is
such that one microphone terminal is connected to the gate of FET, and
the other (which is also connected to the aluminum case of the housing) is
connected to the source of the FET. FET must be provided with a voltage
supply Vdc through a resistor R as shown in the Fig. 3.36 in order to act as
a buffer amplifier. FET is a semiconductor device and acts like a voltage
controlled current source. V (t) is the voltage produced by the microphone
proportional to the sound pressure, and it is the controlling voltage for
the current source. FET converts its voltage input into a current source
output gm V (t), where gm is a parameter of FET, called transconductance.
The terminals of the microphone are these two terminals of the FET. We
will therefore model the microphone output as a current source with a
current output proportional to the voice signal, when appropriately con-
nected to an external circuit.
3.11 Loudspeaker
A loudspeaker is an electro-acoustic transducer [8] that converts electrical energy
into acoustic energy. Unlike electromagnetic energy, acoustic energy requires
a presence of matter in the medium to propagate (acoustic signals doe not
propagate in vacuum). Air is the medium of propagation for audio acoustics,
and the matter that supports the propagation is air. A loudspeaker acts like
a piston and forces air in its vicinity to move at the frequency of the signal
and at an amplitude proportional to the signal amplitude. The structure of an
ordinary loudspeaker is given in Fig. 3.37.
Loudspeakers most commonly have a circular symmetry. The cone section
in Fig. 3.37 is a cone shaped light diaphragm and it simply acts as the piston
head to push the air. It is very lightly supported at the peripheral metal frame
3.11. LOUDSPEAKER 116
by corrugated suspension, both at top and at bottom. This support allows the
diaphragm to move easily, but up and down only.
We need a motor to drive the piston head. The motor is at the lower part.
Diaphragm is rigidly attached to the drive coil. The motor part consists of
a magnetic circuit, which moves the drive coil up and down when there is a
current flowing in the coil. Motor can be analyzed in two parts. The first one
is the magnetic circuit. The magnetic circuit is shown in Fig. 3.38.
The source of the magnetic field is the permanent magnet, whose North-
South poles are aligned vertically in the cross section view. A magnetic flux
emanates from the magnet in that direction as well. The function of the yoke
is to concentrate the magnetic flux into the narrow circular air slit. Yoke is
made of a ferromagnetic material like iron, which conducts the magnetic flux as
copper conducts electric current. Thus almost all the flux (small amount of flux
escapes into surrounding air medium) is concentrated in the slit, generating a
circularly symmetric strong magnetic field, B (top view).
Secondly, a circular drive coil is placed in this field. This is shown in Fig. 3.39.
3.11. LOUDSPEAKER 117
F = N IB (3.71)
where I is the current in the coil, B is the magnetic field in Tesla and N is the
number of turns in the coil.
If the current in the coil is sinusoidal, then the force is obviously sinusoidal.
Whatever the signal (current) is, the force generated is proportional to it. There-
fore, we must apply a current, proportional to the voice signal, to the drive coil
of the loudspeaker. The generated magnetic force is then proportional to the
voice and since the coil is rigidly fixed to the cone membrane (piston), the air
in front of speaker is moved accordingly.
Loudspeakers are specified by their input resistance. 4 Ω, 8 Ω and 16 Ω are
standard input resistance values for this type of loudspeakers.
Headphones are a pair of small loudspeakers worn around the head over a
user’s ears. Earphones are a pair of small loudspeakers that plug into the ear
canal.
3.12 Examples
Example 19
Suppose that the OPAMP shown in Fig. 3.40 is ideal (there is no current into
+ or − terminals, and the output of OPAMP acts like a voltage source.). Find
Vout in terms of Vin .
+0.8v
8K
V2 +
5K Vout
−
V1
Vin
1K 10K
5K
Solution
Since the circuit does not contain any capacitor or inductor, we do not have a
differential equation. Vout can be expressed in terms of Vin algebraically. We
write KCL at node V1 as
V1 − Vin V1 V1 − Vout
. + + =0
1K 5K 10K
or 10(V1 − Vin ) + 2V1 + V1 − Vout =0 We can find V2 from the voltage divider
formula
5K 4
V2 = 0.8 = V
5K + 8K 13
Assume that the output voltage of OPAMP is not saturated and hence it is in
the linear region. From Eq. 3.49, we must have V1 = V2 . Combining equations,
we get Vout = −10Vin + 4. We write the solution as
−10Vin + 4 for Vmin < −10Vin + 4 < Vmax
Vout = Vmax for − 10Vin + 4 > Vmax
Vmin for − 10Vin + 4 < Vmin
Example 20
For the circuit given in Fig. 3.41, assume that the input voltage vin (t) is a step
function, as shown on the right. Assuming Vmin = −15 V and Vmax = 15 V,
find vout (t) for all t > 0.
3.12. EXAMPLES 119
v1 +
1K vout vin(t)
v2 −
vin 1K RL 1V
R1 10K
R2
C t
-v +10nF
C
Solution
This is a first-order RC network. Assume that OPAMP output is not saturated.
We can use the time-domain solution method of Section 2.8 on page 46. (We
cannot use the phasor method since the input is not a sinusoid.)
1. We kill the source vin . In this case, vout = 0. Since v2 = v1 = 0, there is
no current in R1 . Therefore, the total resistance seen by C is R2 only.
2. The time constant is τ = R2 C = 100µs.
3. vout (0− ) = −(R2 /R1 )vin (0− ) = 0 and vC (0− ) = vC (0+ ) = 0V. Hence,
vout (0+ ) = 0V.
4. We open-circuit the capacitor and write
R2
vout (∞) = − vin (∞) = −10V
R1
Since v1 = v2 = 0, vC (∞) = vout (∞) − v2 (∞) = vout (∞) = −10V
5. Since Vmin = −15 < −10, the OPAMP is not saturated and we write vout
as
vout (t) = −10 + (0 − (−10))e−t/τ = −10 + 10e−t/100µ for t > 0
This equation is plotted in Fig. 3.42.
Example 21
For the OPAMP circuit of Fig. 3.41, the input signal is vin (t) = 0.5 cos(2π1000t).
Assuming Vmin = −12 V and Vmax = 12 V, what is vout (t)?
Solution
Assume that the OPAMP output voltage is not saturated. Therefore, the circuit
is linear. Since the input voltage is sinusoidal, we can use phasors. The input
phasor is Vin =0.5. The transfer function is given by
1
Vout R2 ∥ jωC R2 1
=− =−
Vin R1 R1 1 + jωR2 C
3.12. EXAMPLES 120
−1
−2
−3
−4
vout(t) (V)
−5
−6
−7
−8
−9
−10
0 100 200 300 400 500
t (µs)
Vout R2 1 10
= p =p = 8.46
Vin R1 1 + (ωR2 C) 2 1 + (2π103 10−4 )2
Example 22
What is the transfer function, Vout /Vin , of the circuit given in Fig. 3.43, assum-
ing that the input signal is sinusoidal and the OPAMP output is not saturated.
R1 R2 v1 +
Vin
v2 − Vout
C
R3
Solution
We use phasors to find the transfer function. From the voltage divider at the
input side, we find
1
jωC 1
V1 = 1 Vin == Vin
jωC + R1 1 + jωR1 C
R2 is not in the equation since there is no current through it. If the output of
OPAMP is not saturated, we have V1 = V2 . Since there is no current through
R3 , we have Vout = V2 . Therefore, we have
Vout 1
=
Vin 1 + jωR1 C
Example 23
What are the Thévenin and Norton equivalent circuits of the circuit given in
Fig. 3.44(a).
9K 3K 3K 6K 3K 3K
vout vout
+ +
45V − 18K 18K 6K − 30V 18K 6K
(a) (b)
6K 3K 3.6K
vout vout
+ +
20V − 6K 8V −
(c) (d)
Solution
We first find the Thévenin equivalent of 45 V, 9 kΩ and 18 kΩ resistor (inside the
dashed lines): Since 9K∥18K=6K, and 45·18/(18+9)=30 V, we have the equiv-
alent circuit shown in Fig. 3.44(b). One more application gives us the circuit
in Fig. 3.44(c). Finally, we have Req =9K∥6K=3.6K and VT H = 20 · 6K/(6K +
9K) = 8V. We can find the Norton current from IN = VT H /Req =49.4 mA
Example 24
For the circuit in Fig. 3.45, find vout if (a) R2 /R1 =2 and vin =3 V, (b) R2 /R1 =4
and vin =3 V, (c) R2 /R1 =4 and vin =5 V.
3.12. EXAMPLES 122
++15V ++15V
vout
vin − −
R1 -15V R1 -15V
R2 R2
Solution
Both amplifiers are inverting amplifier configurations. We also have Vmin =
−15 V and Vmax = 15 V. The gain of one stage is −R2 /R1 . We have Vmax = 15V
and Vmin = −15V.
(a) vout = (−R2 /R1 )2 vin = 4 · 3 = 12 V
(b) vout = (−R2 /R1 )2 vin = 16 · 3 = 48 V! The second OPAMP is saturated.
vout =15 V.
(c) vout = (−R2 /R1 )2 vin = 16 · 5 = 80 V! Both OPAMPs are saturated.
vout =15 V.
Example 25
Consider the circuit given in Fig. 3.46. Find v3 (t).
560 v3(t)=?
+ R1 R2
v1(t) C L 0.5µ 680 i2(t)
65p
Solution
The linear circuit has two sources at two different frequencies. The phasor
approach can be used to find the solution when all excitations are at the same
frequency. Since we have two different frequencies, we can use the superposition
principle to solve the circuit for one excitation at a time:
When the current source, i2 (t), is killed (open-circuited), the voltage source,
v1 (t), is the only excitation at ω1 = 2π28 · 106 . For this frequency, the capacitor
is replaced with 1/jω1 C, and inductor is replaced with jω1 L. The voltage source
is represented with the phasor V1 = 2∠30o . At this frequency the voltage phasor
V3 is found as
−1
2∠30o
1 1 1
V3 = + + jω1 C + = 1.1∠30o
560 560 680 jω1 L
3.12. EXAMPLES 123
Since the imaginary part of the expression in parentheses is zero, the angle of
the V3 is the same as the angle of V1 . In the time domain, we find
When the voltage source, v1 (t), is killed (short-circuited), the current source,
i2 (t), is the driver at ω2 = 2π60 · 106 . Hence, the capacitor is replaced with
1/jω2 C, and the inductor is replaced with jω2 L. The current source is repre-
sented with the phasor I2 = 13 mA. At this frequency, the voltage phasor V3 is
found as
−1
−3 1 1 1
V3 = 13 · 10 + + jω2 C + = 0.67∠ − 80o
560 680 jω2 L
Example 26
Consider the circuit given in Fig. 3.47(a) with iB (t) = 60 cos(2π103 t) and
(
−5 for t < 0
vA (t) =
+10 for t > 0
Find vC (t).
3K vC(t)=? 1K
+ R1 R2
C 1µ iB(t)
vA(t) R3 2K
(a)
vC(t)=?
(b)
Figure 3.47: (a) Circuit for Example 8, (b) simplified circuit obtained using
Norton equivalent circuits.
3.12. EXAMPLES 124
Solution
The linear circuit has two sources, and the superposition principle is applicable.
The voltage source is a step function at t = 0. Since the circuit is first-order,
we can use the method given on page 46 to find vC (t). On the other hand, the
current source is sinusoidal. For this source, the phasor approach can be easily
applied.
To simplify the circuit, we find the Norton equivalent of the vA (t) and R1
as shown on the left-hand side of Fig. 3.47(b). Similarly, the Norton equivalent
of iB (t), R2 , and R3 are substituted on the right-hand side.
First, we find vC (t) due to current source on the left, while the current source
on the right is killed. vC (0) is found by open-circuiting C for vA (t) = −5 with
t < 0. We get
−5
vC (0) = (3K ∥ 3K) = −2.5
3K
vC (∞) is determined by open-circuiting C while vA (t) = +10 with t > 0:
+10
vC (∞) = (3K ∥ 3K) = 5
3K
The time constant is equal to τ = (1µ)(3K ∥ 3K) = 1.5 ms. Therefore the
solution for this source is
(
−2.5 for t < 0
vC (t) = −t/1.5·10−3
5 + (−2.5 − 5)e for t ≥ 0
We find vC (t) due to the sinusoidal current source on the right, while the
current source on the left is killed. We replace C with 1/jωC with ω = 2π103 ,
the current source with the phasor 2/3 · 60 = 40 mA. The phasor VC is found as
−1
−3 1 1 −0.040
VC = −40 · 10 + + jωC = = 6.33∠96o
3K 3K 0.00632∠84o
Example 27
Consider the 50 Hz high-voltage three-phase circuit given in Fig. 3.48. We have
Va = 380∠0 kV, Vb = 380∠120o kV, Vc = 380∠240o kV all in rms. Find the
rms line currents, I1 , I2 , I3 . Find the total power delivered to load resistances.
Solution
The rms load currents can be found easily since we know the voltages across
the loads:
380K∠0 380∠0
Ia = = = 23.7∠(−5.05o ) A
16K + j1.41K 16.06∠5.05o
3.12. EXAMPLES 125
+ I3
Va 15K 16K
Vc Ic Ia
+
3H 4.5H
Vb I2 Ib
+ I1 4H 14K
380K∠1200 380∠120o
Ib = = = 27.0∠(115o ) A
14K + j1.26K 14.06∠5.13o
380K∠240o 380∠240o
Ic = = = 25.3∠(236o ) A
15K + j0.942K 15.03∠3.60o
The rms line currents can be found from KCL at the nodes:
I1 = Ib − Ia = 27.0∠(115o ) − 23.7∠(−5.05o ) =
= −11.4 + j24.5 − 23.6 + j2.08 = −34.9 + j26.6 = 43.9∠143o A
I2 = Ic − Ib = 25.3∠(236o ) − 27.0∠(115o ) =
= −14.0 − j21.1 + 11.4 − j24.5 = −2.61 − j45.6 = 45.6∠(−93.3o ) A
I3 = Ia − Ic = 23.7∠(−5.05o ) − 25.3∠(236o ) =
= 23.6 − j2.08 + 14.0 + j21.1 = 37.6 + j19.0 = 42.1∠26.8o A
P = |Ia |2 16K + |Ib |2 14K + |Ic |2 15K = 8.95 + 10.2 + 9.6 = 28.8 MW
Example 28
Three OPAMP circuit shown in Fig. 3.49 is known as an instrumentation am-
plifier. Find the output voltage vo in terms of input voltages v1 and v2 .
Solution
Assuming that OPAMPs are not saturated, we have v1 = v5 and v2 = v6 .
Hence ia = (v5 − v6 )/Ra = (v1 − v2 )/Ra . Since the same current flows in the
neighboring resistors, we have
v1 − v2 2R
v3 − v4 = ia (2R + Ra ) = (2R + Ra ) = 1 + (v1 − v2 )
Ra Ra
3.12. EXAMPLES 126
v1 +
v3 R R
1
v5 − ia R
v5 −
ia Ra 3 vo
+
v6 v6
− ia R R R
v2 2 v4
+
3.13 Problems
1. Evaluate the following complex number equations. Find the results in
rectangular and polar format with three significant digits:
(a) (7.82-j10.11)(-2.25-j4.11)
(b) (3.73+j9.32)/(1.28-j2.20)
(c) 3.44∠34o + 2.99∠138o
(d) 1.99 · 10−3 ∠(−27o ) − 8.9 · 10−4 ∠(−150o )
2. Expand ejθ in the Taylor series, group the real and imaginary parts, and
show that the real series corresponds to the expansion of cos θ and the
imaginary series corresponds to the expansion of sin θ.
3. Show that capacitance is a linear circuit element.
4. Show that if a circuit satisfies the linearity definition for two arbitrary
inputs, it also satisfies the linearity condition for an indefinite number of
inputs.
5. A voltage amplifier input/output characteristics is Vo (t) = AVi (t), where
Vi (t) is the input and Vo (t) is the output voltage, and A is a constant
(gain). Show that this amplifier is a linear circuit component.
6. Design an LPF using a resistor and an inductor. Find the transfer function
for this filter and plot its magnitude with respect to angular frequency.
7. Design an HPF using a resistor and an inductor. Find the transfer function
for this filter and plot its magnitude with respect to angular frequency.
8. Find the impedance of the circuits given in Fig. 3.50 at the specified
frequency. Write the impedance in polar form, i.e., magnitude and phase
(2 significant figures):
9. The voltage (current) sources in Fig. 3.51 are connected to the circuits in
problem 8. The frequencies of the sources are as given in problem 8. Find
the current through (voltage across) the sources.
10. Calculate the current through the capacitor in problem 8(c) and inductor
in problem 8(e) and (f) using nodal analysis when the sources in problem 9
are connected across the circuits.
11. The voltage across and the current through two-element series circuits
are given below. Find the component types and their values with two
significant figure accuracy and in regular value notation (like Ω, K for
resistance; µ, p for capacitance, etc.), for each circuit. Determine the
frequency and angular frequency in each case.
(a) v(t) = 28.3 cos(628t + 150o ) V i(t) = 11.3 cos(628t + 140o )
(b) v(t) = 5 cos(2π300t − 25o ) V i(t) = 8 cos 2π300t + 5o ) mA
(c) v(t) = 10 cos(2π796t−150o ) V i(t) = 1.333 cos(2π796t−3π/8) mA
6 o
(d) v(t) = 8 cos(10 t + 45 ) V i(t) = 8 cos(106 t + 90o ) mA
(e) v(t) = 5 cos(2π106 t − 160o ) V i(t) = 10 cos(2π106 t − 75o ) mA
3.13. PROBLEMS 129
12. A series circuit has a resistor R=120 Ω and an inductor L=780 nH. A
voltage of 10 V peak value with a frequency of 25 MHz (zero phase)
is applied across this circuit. Find the current flowing through it and
write down the expression for the time waveform. Find the current, if the
frequency is increased to 50 MHz.
13. A series circuit has R=1 kΩ and C=120 pF. What is the frequency (not
angular frequency) at which the phase difference between the current and
voltage is π/4?
14. A series RC circuit has C=470 pF. Find R if the phase difference between
current and voltage is 30o at 1 kHz.
15. The voltage and current of a two-element series circuit at 500 kHz are
V =3∠45o V and I=1∠120o mA. When the frequency is changed to another
value, f , the phase difference between the voltage and current becomes
30o . Find f .
16. Assume that the voltage and current pairs given in problem 11 are for
two-element parallel circuits. Determine the component types and their
values.
17. Find and draw the Thévenin equivalent of the circuits given in Figs. 3.52
and 3.53. Express the equivalent voltages and impedances in polar form.
18. Convert the equivalent circuits found in problem 17 into Norton equivalent
circuits and draw them.
3.13. PROBLEMS 130
19. Find and draw the Norton equivalent of the circuits given in Fig. 3.54.
Express the equivalent currents and impedances in polar form.
20. Find the Veq and Zeq such that the circuit given in Fig. 3.55(a) can be
represented as in (b).
21. Find and draw the Thévenin and Norton equivalents of the circuits in
Fig. 3.56 at DC, 20 MHz, and 40 MHz.
22. This problem illustrates how a unity feedback amplifier is used to avoid
loading effects. Consider the divider circuit in the Fig. 3.57(a). What is
Vout ? Assume we want to apply Vout across a 1 kΩ resistor, as shown
in part (b). What is Vout now? Now assume we place a buffer amplifier
between the divider and 1K resistor as in part (c). Find Vout .
3.13. PROBLEMS 131
23. Assume there are two signals, V1 and V2 . Design a summing amplifier to
produce Vout = 2V1 +0.5V2 , using (a) two OPAMPs, and (b) one OPAMP.
assuming that OPAMPs are ideal.
24. Make a table indicating what terminals to connect to the input signal
source or the output to get all possible (different) amplification factors for
the circuit in Fig. 3.58. Also, calculate the resulting input impedance and
the possible gains, and include them in the table.
25. Find Vo /Vin in the OPAMP circuits of Figs. 3.59 and 3.60.
26. Find the transfer functions Vo (ω)/Vin (ω) for the circuits given in Figs. 3.61
and 3.62.
27. In the circuits of Fig. 3.61 and 3.62, find the asymptotic behavior of the
transfer function at low frequencies and high frequencies.
28. Find the transfer function of the circuit given in Fig. 3.63. Is there a
frequency at which the gain is zero? Which frequency?
29. Find the voltage output of a condenser microphone of capacitance Co
as a function of diaphragm vibration when a cable of capacitance Cc is
connected at its output.
3.13. PROBLEMS 132
DIODES and
BIPOLAR JUNCTION
TRANSISTORS
Diodes are important electronic devices used in many applications. Unlike the
devices like resistors, capacitors, and inductors, they are nonlinear devices. In
the first part of this chapter, we explore the characteristics of diodes as a circuit
element. We study the solution of circuits containing diodes. Several important
circuit containing didoes are presented.
As seen in the previous chapter, OPAMPs can be used for low-frequency (or
audio frequency) amplification. At high frequencies, discrete transistors or in-
tegrated circuit amplifiers are used for amplification. At microwave frequencies,
MMICs (monolithic microwave integrated circuits) are preferred for the same
purpose. In the second part of this chapter, we explore the bipolar junction
transistors. After introducing the DC solution of circuits containing a BJT, the
amplification property of the circuit is demonstrated.
4.1 Diodes
A diode is a nonlinear resistor with a symbol shown in Fig. 4.1(a). It carries the
current in one direction but not in the other. An ideal diode is described with
two states: OFF state—it carries no current with a negative voltage across it;
ON state—it has no voltage across it with a positive current through it.
ON state: vD = 0 if iD ≥ 0
OFF state: iD = 0 if vD < 0 (4.1)
iD iD iD iD
+ +
vD vD
- - + =0.7V
V0
vD vD
0.7
Figure 4.1: (a) Diode symbol and reference directions, (b) I − V characteristics
for an ideal diode, (c) Approximate diode equivalent circuit, (d) I − V charac-
teristics for the diode approximate equivalent circuit
Real diodes
Most of the contemporary diodes are semiconductor devices. They are built by
a junction of p-type and n-type semiconductors called p-n junction diodes. They
are nonlinear resistors, resistance of which depend on the voltage across them.
The I − V characteristic of a semiconductor diode can be well approximated by
an exponential relation:
ID = IS (e−VD /γ − 1) (4.2)
where ID and VD are the current and voltage of the diode. IS and γ are the
physical constants related to the material and construction of the diode. The
typical I-V characteristic of a silicon p-n junction power diode is given in Fig. 4.3.
ON state: vD = Vo V if iD ≥ 0
OFF state: iD = 0 if vD < Vo V (4.3)
4.1. DIODES 136
anode iD iD (A)
+ 3
vD
- 2
cathode 1
-400
1 vD (v)
breakdown
voltage
The procedure for the solution of circuits containing one or more diodes
is as follows:
1. Assign ON or OFF state to the diode(s).
2. Replace the diode(s) with the equivalent circuit in that state.
4. Check if the condition of the assumed state is satisfied for the diode(s).
If it is difficult to estimate the state of the diode, it is better to try the OFF
state first, since it usually results in a simpler circuit. If there are several diodes,
the procedure should be repeated until the conditions of all diodes are satisfied.
The procedure is best understood by the examples given below.
Example 29
V1 Va
iR iR
4K 3K 4K 3K 4K 3K
D
3m 3m V2 +vD - V3 3m Vb + Vc
0.7V
5K 2K 5K 2K 5K 2K
We want to find the current iR in the circuit given in Fig. 4.5 with one diode.
Use the approximate model of the diode in Eq. 4.3. Apply the procedure:
1. Assume that the diode is OFF.
2. Replace the diode with an open-circuit (as in Fig. 4.5(b)).
3. Solve the resulting linear circuit:
6. Since the diode voltage vD =1.5 V>0.7 V, the OFF condition of Eq. 4.3 is
not satisfied: The diode must be ON.
4.1. DIODES 138
Example 30
10K 4K D 10K 4K D
+ iR vC(0)=-20V + + iR +
50V vC + 2µF 5V 50V vC + 2µF 5V
(a) (b)
10K 4K D 10K 4K D
+ iR + + iR +
50V vC + 2µF 5V 50V v+
C
5V
(c) (d)
We want to find the current iR in the circuit given in Fig. 4.6 with one ideal
diode (V0 =0). Use the approximate model of the diode in Eq. 4.3. Apply the
procedure:
1. Assume that the diode is OFF.
2. Replace the diode with an open-circuit (as in Fig. 4.6(b)).
3. Solve the resulting linear circuit:
and for iR as
6. We check that for t >8.84 m, the diode current (=iR ) is positive, so our
assumption of the diode being ON is correct.
A MATLAB code to plot vC and iR is given below. The resulting plot is shown
in Fig. 4.7.
% MATLAB code to draw iR and vC of Example 8
clear all hold off
t1=0:0.01:8.84; % define two separate vectors
t2=8.84:0.01:20; % for two regions
iR1=7*exp(-t1/20); % current for the first region
iR2=3.21+(4.5-3.21)*exp(-(t2-8.84)/5.72); % second region
4.1. DIODES 140
Example 8
7 20
6 10
iR
vC
i R (mA)
v C (V)
5 0
4 -10
3 -20
0 2 4 6 8 10 12 14 16 18 20
t (ms)
(a) (b)
Vp
Vp -V0
Vp /2
Voltage (volts)
vin (t)
-Vp /2
vo(t)
-Vp
0 T/4 T/2 3T/4 T 5T/4 3T/2 7T/4 2T 9T/4
time
Figure 4.8: (a) Half-wave diode rectifier, (b) the equivalent circuit, (c) the input
AC voltage, vin (t), with a peak value, Vp ; the output voltage, vo (t), on the load
resistor, R.
at the input, while it becomes zero during negative half cycles. The current
starts flowing as soon as vin exceeds V0 = 0.7V and stops when vin falls below
V0 . The voltage that appears across the load, vo , is, therefore, sine wave halves
as depicted in Fig. 4.8(c). This voltage waveform is neither an AC voltage nor
a DC voltage, but it is always positive.
When this circuit is modified by adding a capacitor in parallel with R, we
obtain the circuit in Fig. 4.9(a), and its equivalent Fig. 4.9(b). The capacitor
functions like a filter together with the resistor, to smooth out vo (t).
Consider the vo (t) waveform in Fig. 4.9(c) with RC = 10T .
When vin (t) − V0 exceeds vo (t) (at t=tr ), the diode starts conducting, and
the current through the diode charges up the capacitor.
After the peak voltage is reached, the voltage at the anode of the diode,
vin , falls below Vp and hence the voltage across the diode, vin −vo , becomes
less than Vo . The current through the diode ceases flowing. The diode is
reverse biased.
Now, the AC voltage source is isolated from the parallel RC circuit, and
the capacitor is charged up to Vp −Vo . The capacitor starts discharging on
R with a time constant of RC. If RC is small, as depicted in Fig. 4.9(c)
for RC = T , the capacitor discharges quickly. If RC is large, the discharge
is slow, as shown in the case with RC = 10T .
4.1. DIODES 142
ideal
p-n diode diode 0.7V
+
+ + + +
vin(t) C R vo(t) ~
~ vin(t) C R vo(t)
- -
(a) (b)
Vp
Vp -V0
Vp /2
Voltage (volts)
v in(t)
v o (t) with C=0
-Vp /2
v o (t) with RC=T
v o (t) with RC=10T
-Vp
0 T/4 T/2 3T/4 tr T 5T/4 3T/2 7T/4 T+tr 2T 9T/4
time
Figure 4.9: (a) Half-wave diode rectifier with RC filter, (b) equivalent circuit
and (c) voltage waveforms on load resistor for RC = T and RC = 10T .
As vin increases for the next half-cycle of positive sine wave tip, it exceeds
the voltage level to which the capacitor discharged until then, at t=T + tr ,
and the diode is switched on again.
It starts conducting, and the capacitor is charged up to Vp − V0 all over
again (t=2T ).
The fluctuation in the output voltage is called ripple. Clearly, the case for
RC = T has a larger ripple than the case for RC = 10T .
Example 31
Find the peak-to-peak ripple voltage amplitude for the half-wave rectifier when
vin (t) = 25 cos(2π50t), V0 =0.7 V, R = 470 Ω and C = 1000 µF.
We have T = 1/50Hz=20 ms. τ = RC = 470 · 1000 · 10−6 = 470ms. Since
T ≪ τ , we can use Eq. 4.7:
T 20
Vr ≈ (Vp − V0 ) = (25 − 0.7) = 1.0 Vpp
τ 470
This circuit is also solved with LTSpice circuit simulator (see page 300 for a
tutorial on LTSpice). The capacitor voltage and diode current obtained after a
transient simulation are given in Fig. 4.10. The results show that the peak-to-
peak ripple voltage is 0.95 V (our estimate of 1 V is reasonably close). The peak
diode current is 1.68 A, while the average resistor current is only 50.5 mA. Note
that the diode current exists only during a short period while the capacitor is
being charged.
4.1. DIODES 144
Figure 4.10: LTSpice schematic of the half-wave rectifier and the results of
LTSpice simulation for the output voltage, the current in the diode, and the
resistor.
Newton-Raphson method
We can use the Newton-Raphson method (named after English Mathematician
and Physicist Isaac Newton (1642–1727) and English Mathematician Joseph
Raphson (1648–1725)) the find the zero of a function f (x) in an iterative manner:
f (xn )
xn+1 = xn − (4.9)
f ′ (xn )
where f ′ (x) is the derivative of the function f (x). We stop the iterations, when
xn+1 and xn are very close to each other.
Example 32
Find the peak-to-peak ripple voltage amplitude for the previous example when
R=220 Ω and C = 100µF.
τ = RC = 220 · 100 · 10−6 = 22ms. Since T ≪ τ is not satisfied, we must
find the intersection point, tr , numerically from
2π
vo (tr ) = (Vp − V0 )e−tr /τ = Vp cos( tr ) − V0
T
or
2π
24.3e−tr /τ = 25 cos( tr ) − 0.7
T
To find the value of tr , let us use Newton-Raphson method to find the zero of
the function
2π
f (t) = 24.3e−t/τ − 25 cos( t) + 0.7 = 0
T
4.1. DIODES 145
24.3e−t/τ − 25 cos( 2π
T t) + 0.7
tn+1 = tn − 24.3 −t/τ
− τ e + 25 T sin( 2π
2π
T t)
ripple=(Vp-V0)*(1-exp(-t/tau)) % result
rippleEst=(Vp-V0)*T/tau % approx result
In the negative half-cycle, the lower AC source provides the current since
D2 is conducting, and D1 is OFF.
As shown in Fig. 4.11(b), the capacitor is charged in the same direction to the
same peak value: Vp − V0 . Thus, the capacitor is charged to the peak value
twice in one cycle, reducing the amplitude of ripple.
+ τ= RC vo(t)
vAC + D1 +
vo(t) Vp-V0
C R
Vp cos(ωt) -
vAC +
Vp cos(ωt) D2 tr
T/2 T
(a) (b)
Figure 4.11: (a) A full-wave rectifier utilizing two identical AC sources, (b) the
voltage vo (t) across the capacitor.
T /2 T
Vr ≈ (Vp − V0 ) = (Vp − V0 ) if T /2 ≪ τ (4.10)
τ 2RC
D1
+vo(t) τ= RC vo(t)
vAC + + Vp-V0
D3 C1 R1
Vp cos(ωt)
vAC + +
D2 R2
C2 tr
Vp cos(ωt) -vo(t)
T/2 T
D4
(a) (b)
Figure 4.12: (a) Two full-wave rectifiers utilizing two identical AC sources, (b)
the voltage ±vo (t) across the load resistors.
When vAC is in its positive phase, D2 and D4 conduct, and current flows
through D2 , the capacitor, and D4 until the capacitor is charged up to
the peak value, Vp − 2 V0 . The peak voltage for vL is less than the one
in a single diode case because the charging voltage has to overcome the
threshold voltage of two diodes instead of one.
During the negative half-cycles, D1 and D3 conduct, and the capacitor
is thus charged up in the negative phase as well. Since the capacitor is
charged twice in one cycle of vAC the ripple in the waveform of Fig. 4.13(c)
is nearly the same as that in the full-wave rectifier.
D1 τ= RC vo(t)
D2 +
vAC + + Vp-2V0
vo(t)
C R
Vp cos(ωt) D4 D3 -
tr
T/2 T
(a) (b)
Figure 4.13: (a) Bridge rectifier, (b) rectified output voltage without capacitor
(thin curve), and filtered output voltage (thick curve).
* The same four-diode bridge configuration can also be used for the two full-wave rectifiers
of Fig. 4.12(a).
4.1. DIODES 148
T /2 T
Vr ≈ (Vp − 2V0 ) = (Vp − 2V0 ) if T /2 ≪ τ (4.11)
τ 2RC
Example 33
Find the peak-to-peak ripple voltage amplitude for the bridge rectifier when
vAC (t) = 25 cos(2π50t), R = 470Ω and C = 1000µF.
From Eq. 4.11:
T /2 10
Vr ≈ (Vp − 2V0 ) = (25 − 1.4) = 0.5Vpp
τ 470
The ripple voltage is half of the ripple for the half-wave rectifier circuit of p. 143
with the same R and C values.
anode iD iD (mA)
+ 30
vD
- 20
cathode
10
-VZ
-5.6 0.7 vD (v)
zener -10
breakdown
voltage -20
-30
When a zener diode is used in a circuit given in Fig. 4.15(a), a reverse diode
current
VS − VZ
I = −ID = (4.13)
R
, flows through the diode as long as VS > VZ . Vo = VZ appears across the diode
independent of the value of VS as long as VS > VZ . We call this action “voltage
regulation”. On the other hand, if VS is less than VZ , the diode is no longer in
the breakdown region, and behaves like an open circuit. In that case, we have
Vo = VS , and no voltage regulation.
+VS +VS
I R I R
Vo Vo
IL RL
(a) (b)
Assume that a load resistor RL is connected across the zener diode, as shown
in Fig. 4.15(b). If the zener diode is in the breakdown region and a reverse
current flows, we have Vo = VZ . In this case, the current through RL is
VZ
IL = (4.14)
RL
To satisfy this condition, we must have
VS − VZ RL
I= > IL or VZ < VS (4.15)
R R + RL
Therefore, Vo = VZ is independent of the value of VS , and voltage regulation is
achieved and as long as Eq. 4.15 is satisfied. We note that under this condition,
the zener diode dissipates a power of
VS − VZ VZ
PZ = VZ (I − IL ) = VZ − (4.16)
R RL
which must be less than the allowed power dissipation rating of the zener diode.
4.1. DIODES 150
If the condition in Eq. 4.15 is not satisfied, the zener diode remains off. The
output voltage is determined by the voltage divider formed by R and RL :
RL
Vo = VS (4.17)
R + RL
and there is no voltage regulation.
spring
forward direction reverse direction reverse direction
small pressure small pressure large pressure
ON OFF ON
(a) (b) (c)
Figure 4.16: Water flow analogy of a zener diode: (a) Current/water flows in
the forward direction, (b) current/water does not flow in the reverse direction,
(c) current/water flows in the reverse direction with sufficient voltage/pressure.
Example 34
Design a 5.6 V zener diode regulator circuit for a voltage source VS , which varies
between 10 V to 13 V (for example, due to ripple). Suppose that we have a 5.6 V
zener diode, which can at most dissipate PZmax =300 mW. Find the smallest
load resistor while the regulation is still performed. For a good regulation the
minimum zener current should be 1 mA.
The maximum current that the zener diode can carry is found from its power
dissipation limit:
PZmax 0.3W
−ID = = = 0.053 A = 53 mA
VZ 5.6V
Referring to Fig. 4.15, we choose the value of R using the no-load condition
(i.e., when there is no RL ) and under the maximum VS .
VSmax − VZ 13 − 5.6
R= = ≈ 140 Ω
−ID 0.053
Let us choose the next largest standard resistor value: R=150 Ω. With this
R, the worst-case load current is supplied when VSmin =10 V is applied. The
4.1. DIODES 151
current flowing in R is
VSmin − VZ 10 − 5.6
= = 29 mA
R 150
Since we must reserve at least 1 mA for the zener diode itself, we have 28 mA
remaining for the load current. This current is only sufficient for a load resistor
of
5.6
RL = = 210 Ω
0.028
Hence the zener diode regulator supplies a constant 5.6 V output voltage for
load resistors in the range 210 Ω < RL < ∞, while the input voltage VS varies
between 10 to 13 V. With any VS and RL in this range, the zener diode current
is such that the output voltage is 5.6 V. Any excess current coming from the
source side through R flows in the zener diode, and the zener diode dissipation
is less than 300 mW limit.
Zener diodes are useful as voltage regulators only when the load current
demand is low. For higher current needs, integrated-circuit voltage regulators
must be utilized.
4.1.8 LED
Light-emitting-diode (LED) is a special semiconductor diode that emits light
when a current flows. The symbol and package polarity of an LED are shown in
Fig. 4.17(a) and (b). Since it is a diode, it carries current only in one direction.
The color of the light is determined by the type of semiconductor used in the
fabrication of the diode. The energy bandgap of the semiconductor determines
the wavelength of the emitted photons. LEDs exist in different colors of the
visible spectrum as well as infrared and ultraviolet regions.
The voltage drop across an LED is higher than 0.7 V of a regular silicon
diode. Red-colored LED’s have a voltage drop of 1.6 to 1.7 V. The voltage drops
of the infrared LEDs are lower while those of green LEDs are higher. Blue LEDs
have even higher voltage drops. The current-voltage characteristic of an LED
is similar to that of a silicon diode. The current increases exponentially as a
function of voltage, meaning that a small change in voltage can cause a large
change in current. Not to exceed the maximum current rating of LEDs, they
are typically driven by a resistor in series that limits the current. Referring to
+ R
+
Vs Id
-
(a) (b) (c)
Figure 4.17: (a) LED Symbol, (b) LED package showing the polarity, (c) LED
drive circuit.
4.2. BIPOLAR JUNCTION TRANSISTOR (BJT) 152
Vs − Vo
R= (4.18)
Id
where Id is the desired current through the LED and Vo is the voltage drop
across the LED (Vo ≈ 1.6 V for a red LED).
The light conversion efficiency of LEDs is higher than that of incandescent
lamps. Modern power LEDs are being used for illumination purposes, replacing
the conventional lamps. A typical LED lamp contains many LEDs connected
in series to increase the light output. The schematic of a low-power LED lamp
is given in Fig. 4.18. The current flowing through the LEDs is in the 30 mA to
200 mA range. The voltage to drive the LED is obtained by a bridge rectifier
and an electrolytic capacitor C2 . To reduce the 220 Vrms line voltage to the
lower voltage necessary to drive the LEDs, a high-voltage capacitor, C1 , is used
in series with the bridge rectifier. Since a capacitor does not dissipate any
energy, it is a low-cost and energy-saving solution to reduce the voltage. A
small resistor in series is used to limit the current and as a protection to reduce
the risk of fire in case the series capacitor fails.
C1
220Vrms
C2 +
R
John Bardeen in 1947. A BJT is essentially two p-n junction diodes connected
back-to-back, sharing either p or n regions. If the diodes share the p region,
the resulting BJT is called NPN type. If the n-region is the shared region, the
resulting BJT is of PNP type. The shared region of the diodes is called the
base. The other two terminals of a BJT are called emitter and collector.
The symbols of the two types of BJTs are shown in Fig. 4.19. The arrow in
the symbol is always in the emitter terminal and indicates the direction of the
emitter current, IE .
collector collector
NPN IC PNP IC
base base
IB IE IB IE
emitter emitter
In the cutoff state, no current flows in the emitter-base junction (IB =0) and
no current flows in the collector-base junction (IC =0). The collector acts like
an open-circuit. In the active state, we have the transistor action and high
current gain results. In this state, Eqs. ?? and ?? are valid. The collector acts
like a current-controlled current source. In the reverse active mode, the roles
of emitter and collector are interchanged, and a low current gain results. In
the saturation mode, the collector current is not determined by Eq. ??. Hence
Eqs. ?? and ?? are not valid. In that case, the collector acts like a voltage
source of value VSat . and external circuitry determines the collector current.
Since the current gain is much smaller, the reverse-active region is not a
preferred operation mode, and it should be avoided except in very rare circum-
stances. In this text, from this point on, the reverse-active state will not be
dealt with.
We consider only the three states, cutoff (OFF), active (ACT), and satura-
tion (SAT) states, as summarized in Table 4.2 for an NPN BJT.
In this table, V0 is the turn-on voltage of the base-emitter diode, and it is
in the range 0.6 to 0.8 V. VSat is the collector-emitter saturation voltage, and
it is in the range 0.1 to 0.3 V for small BJTs.
The states of a PNP BJT is similar as shown in Table 4.3. The polarities of
base-emitter and collector-emitter voltage are changed.
4.3. DC ANALYSIS OF BJT CIRCUITS 154
We note that the current gain β of a BJT varies in a wide range (e.g., 180 to
460) even for the same brand and the same model transistor. β is also dependent
on the value of the collector current, temperature, and age of the BJT.
b c b c b c
IB I IB IC
+ C + +
V0 V0 VSat
βIB
e e e
OFF ACT SAT
Simple DC models for different states of a PNP BJT are given in Fig. 4.21.
The procedure to find the state of a NPN BJT is similar to the procedure
to find the state of diodes as described earlier:
b c b c b c
IB I IB IC
C
V0 V0 V
βIB + + Sat
+
e e e
OFF ACT SAT
The procedure for a PNP transistor is similar, with some quantities changing
sign:
Example 1
Find the range of values of RC such that the NPN BJT in the circuit of Fig. 4.22
stays in ACT region. We have VCC =8 V, V0 =0.7 V, RB =220 KΩ, VSat =0.2 V
and β=120.
Solution
Assume that the NPN BJT is in the ACT state. We find IB =(VCC −V0 )/RB =(8−
0.7)/220=0.033 mA. Hence IC =βIB =3.98 mA. To be in the active region, we
4.4. BIASING OF BJTS 156
VCC
IC
RB RC
IB +
+ VCE
V0 -
-
must have VCE =VCC − RC IC > VSat . Hence RC < (VCC − VSat )/IC =1.95 KΩ.
Example 2
Find the range of values of RB such that the PNP BJT in the circuit of Fig. 4.23
stays in ACT region. We have VCC =6 V, V0 =0.7 V, RC =1.8 KΩ, VSat =0.2 V
and β is in the range 80 to 130.
+ VCC
IB V0 -
-
VCE
+
RB IC
RC
Solution
Assume that the PNP BJT is in the ACT state. We must have −VCE >
VSat =0.2 V. Since −VCE =VEC =VCC −RC IC =6−1.8IC >0.2 V. Hence IC <3.22 mA.
For the worst case, we use β=130. IB <3.22/130=0.025 mA. Since IB =(VCC −
V0 )/RB =(6 − 0.7)/RB <0.025 mA. We find RB > 214 KΩ. Note that a smaller
value of β will result in a smaller IC , and hence guaranteeing the active state
condition.
VCC
IC
RB RC
IB +
+ VCE
V0 -
-
determined from
VCC − Vo
IB = (4.19)
RB
To analyze the circuit, we assume that the BJT is in ACT region. So we have
IC = βIB . The collector voltage is determined by the the collector current and
the collector resistance, RC :
where β is the current gain of the BJT. To check our assumption of BJT being
in ACT region, we must check that VCE > VSat .
Otherwise, the BJT is in SAT region, and we have VCE = VSat and
VCC − VSat
IC = (4.21)
RC
Example 3
Let us consider the schematic in Fig. 4.24 with RB =470 KΩ, RC =1 KΩ, VCC =12 V,
V0 =0.7 V, and VSat =0.2 V, while β varies from 180 to 460 device to device for
a particular transistor.
We assume that the BJT is in ACT state. With β=180 from Eq. 4.19 we find
IB =24 µA, from Eq. ?? IC =4.3 mA, and from Eq. 4.20, we have VCE =7.7 V.
Since 7.7 > 0.2 V our ACT state assumption is valid.
For β=460, we find IC =11 mA, VCE = 0.94 V. Since 0.94 > 0.7, the BJT is
(luckily) still in the ACT state.
Example 4
Let us consider the same circuit above with RC =1.2 KΩ.
4.4. BIASING OF BJTS 158
We assume that the BJT is in ACT state. With β=180 from Eq. 4.20, we
have VCE =6.8 V. Since 6.8 > 0.2 V our ACT state assumption is valid.
For β=460, we find VCE = −1.3 V. Since −1.3 < 0.7, the BJT is not in the
ACT state. It is in the SAT state. We have VCE = VSat = 0.2 V and from
Eq. 4.21, IC =9.8 mA̸= βIB .
VCC
IC
RB RC
IB +
+ VCE
V0 -
-
IE RE
Figure 4.25: Biasing an NPN transistor using a base resistor and an emitter
resistor.
VCC − V0
IB = (4.22)
RB + (β + 1)RE
Again we need VCE > VSat for the verification of BJT ACT state operation.
If the transistor is found to be in SAT state, then we have VCE = VSat and
use nodal analysis to find the emitter voltage:
VE VCC − V0 − VE VCC − VSat − VE
− − =0 (4.24)
RE RB RC
We can determine the collector current from
VCC − VSat − VE
IC = (4.25)
RC
4.4. BIASING OF BJTS 159
Example 5
Let us consider the schematic in Fig. 4.25 with RB =560 KΩ, RC =1 KΩ, RE =470 Ω,
VCC =12 V, V0 =0.7 V, and VSat =0.2 V while β=180–460.
First, we assume the transistor to be ACT. From Eq. 4.22, we have IB =17 µA
(for β=180) or IB =14 µA (for β=460). From Eq. 4.23 we have VCE =7.4 V (for
β=180) or VCE =2.1 V (for β=460). The variation in VCE is smaller in this case.
VCC VCC
IC IC
RB1 RC RC
IB + IB +
RT
+ VCE + VCE
V0 - + V0 -
- -
VT
RB2 IE RE IE RE
(a) (b)
equivalent circuit of the two base resistors and the supply voltage as shown in
Fig. 4.26:
RB2 RB1 RB2
VT = VCC and RT = (4.26)
RB1 + RB2 RB1 + RB2
We assume ACT state for BJT and using KVL, we can find the base current as
VT − V0
IB = (4.27)
RT + (β + 1)RE
Example 6
Let us consider the schematic in Fig. 4.26(a) with RB1 =18 KΩ, RB2 =4.7 KΩ,
RC =1 KΩ, RE =470 Ω, VCC =12 V, V0 =0.7 V, and VSat =0.2 V while β=180–
460. Find the range of VCE values.
4.4. BIASING OF BJTS 160
Solution
First, we assume the transistor to be ACT. From Eq. 4.26, we have VT =2.5 V
and RT =3.7 KΩ. From Eq. 4.27, we get IB =20 µA (for β=180) or IB =8.1 µA
(for β=460). From Eq. 4.23 we have VCE =6.6 V (for β=180) or VCE =6.5 V
(for β=460). Since VCE > VSat =0.2 V, BJT is ACT for the whole range of β.
Clearly, the variation in VCE is much smaller with this biasing circuit.
Example 7
Let us consider the PNP BJT in the schematic of Fig. 4.27(a) with RB1 =33 KΩ,
RB2 =8.2 KΩ, RC =2.2 KΩ, RE =680 Ω, VCC =15 V, V0 =0.7 V, and VSat =0.2 V
while β=90–180. Find the range of VCE .
VCC VCC
IE RE IE RE
RB2
+ +
IB V- 0 - RT IB V0- -
VCE VCE
+ +
+
RB1 VT
IC RC IC RC
(a) (b)
Solution
First, we assume the transistor to be ACT. We have VT =(RB1 /(RB1 +RB2 )VCC =12 V
and RT =RB1 ∥ RB2 =6.57 KΩ. From
VCC − VT − V0
IB =
RT + (β + 1)RE
For β=90, IB =0.0336 mA, IC =3.02 mA, and for β=180, IB =0.0177 mA, IC =3.19 mA.
Since −VCE = VCC − RE IE − RC IC , we find IE =3.05 mA, and −VCE =5.90 V
for β=90; IE =3.21 mA, and −VCE =5.79 V. Since −VCE > VSat =0.2 V, the
PNP BJT is ACT for the whole range of β.
VCC
RC
RB1 A
IC
IB +
B VCE
+
RB2 V0 -
-
Figure 4.28: Biasing an NPN transistor using two base resistors with collector
feedback.
we assume that the BJT is ACT state and write the nodal equations at nodes
A and B:
VCE − VCC VCE − V0
+ IC + =0 (4.28)
RC RB1
V0 IC V0 − VCE
+ + =0 (4.29)
RB2 β RB1
since IB =IC /β. Solving for IC and substituting in the other equation, we get
β+1 1 β+1 β VCC
+ VCE = + V0 + (4.30)
RB1 RC RB1 RB2 RC
to determine VCE . There is no need to check that VCE > VSat , since VCE >
V0 > VSat at all times. So, this biasing arrangement can never go wrong as
long as RB2 is not selected too small. Indeed, if we ignore the small terms in
Eq. 4.30 we find
RB1 + RB2 RB1
VCE ≈ V0 if RC > (4.31)
RB2 β
We note that the BJT in this biasing arrangement can never be in the SAT
state since VCE > V0 to get a positive IB . On the other hand, one should make
sure that
RB1 + RB2
V0 < VCC (4.32)
RB2
otherwise, the transistor is in the OFF state.
Example 8
Consider the schematic in Fig. 4.28 with RB1 =33 KΩ, RB2 =4.7 KΩ, RC =1 KΩ,
VCC =12 V, and V0 =0.7 V while β=180–460.
From Eq. 4.30, we have VCE =6.6 V (for β=180) or VCE =6.0 V (for β=460).
The variation in VCE is small in spite of the large variation in β.
4.5. SMALL-SIGNAL BJT MODEL 162
b c b c b c
ic
rbe ib βib rbe ib ic
e e e
OFF ACT SAT
Figure 4.29: Small-signal model of a BJT (NPN or PNP) in the OFF, ACT,
and SAT states.
kT
rbe = (4.33)
qIB
where k is the Boltzmann constant (k = 1.38 × 10−23 J/◦ K), T is the tem-
perature of the B-E junction in Kelvins, q is the charge of an electron (q =
1.6 × 10−19 coul), and IB is the DC B-E junction current in Amperes. As the
base current, IB , gets larger, the small signal resistance, rbe becomes smaller.
At room temperature (T = 300◦ K) we have
kT
≈ 0.0259 V (4.34)
q
Therefore
0.0259 25.9
rbe (Ω) ≈ = (4.35)
IB (A) IB (mA)
The small-signal analysis of a NPN (or PNP) BJT circuit in the ACT state
can be summarized as follows:
3. If IB > 0 and VCE > VSat (−VCE > VSat for PNP), the transistor is
in the ACT (active) state. In that case, substitute the small-signal
model of the transistor. Use the DC base current, IB , in calculating
rbe value.
4. Keep the AC small-signal sources, and kill the DC sources.
5. If there are capacitors whose reactances are much smaller than the
resistors in the circuit, they can be shorted. If there are inductors
whose reactances are much larger than the resistors in the circuit,
they can be open-circuited.
6. Perform a small-signal AC analysis of the circuit. If you have capaci-
tors or inductors left, use their phasor equivalents.
Example 9
Consider the BJT amplifier shown in Fig. 4.30(a) with VCC =12 V, V0 =0.7 V,
RB =470 KΩ, and RC =1 KΩ with β=180–420. From the DC analysis of the
circuit, we know that the BJT is in the ACT state for all β values. Hence we
can use the small-signal model of Fig. 4.29. Assuming that DC-block capacitors,
C1 and C2 , are sufficiently large at the frequency of the input sinusoidal source,
find the small-signal voltage gain of this amplifier (vout /vin ). The DC-block
capacitors are there to prevent the bias circuit to be affected by the input
voltage source or output load that may be present.
VCC
IC
RB RC ib
C1 IB vout + vout
+ C2
+ VCE vin RB rbe βib RC
+
vin V0 -
-
(a) (b)
Figure 4.30: (a) A simple BJT amplifier, (b) The small-signal model of the
amplifier.
4.5. SMALL-SIGNAL BJT MODEL 164
Note that the negative sign in front of the gain value signifies that there is a
180o phase change from input to output. As β varies from device to device, the
voltage gain will also vary in this case.
Example 10
Let us find the small-signal voltage gain of the amplifier shown in Fig. 4.31(a)
with RB1 =33 KΩ, RB2 =4.7 KΩ, RC =1 KΩ, VCC =12 V, V0 =0.7 V, and β=180–
460. We know that the BJT is in ACT state, so we can use the small-signal
model of Fig. 4.29. DC-block capacitors, C1 and C2 are sufficiently large to act
like short-circuits at the operating frequency.
From the DC analysis of the circuit, we have IB =0.029 mA, rbe =0.89 K
(for β=180) and IB =0.013 mA, rbe =1.99 K (for β=460). We write the node
equation at node A as
vout vout − vin
+ + βib = 0 (4.41)
RC RB1
We also have
vin
ib = (4.42)
rbe
Hence we find the voltage gain as
VCC
RC
RB1 A
vout RB1
C2 B A vout
C1 B IB + +
+ + VCE vin RB2 rbe βib RC
vin RB2 V0 -
-
(a)
Figure 4.31: (a) A BJT amplifier with collector feedback, (b) The small-signal
model of the amplifier.
or
−196 for β = 180
Av = (4.44)
−224 for β = 460
Note that the gain variation is small in spite of the large variation in β.
We note that if the BJT is in SAT or OFF state, no voltage gain should be
expected. Therefore, a good DC biasing design is essential to get a small-signal
voltage gain from a BJT.
Example 11
Find the gain of the circuit given in Fig. 4.32(a) with RB =10 K, RC =1 K, and
VCC =12 V.
VCC VCC
RC RC
A A
vout vout
RB
C1 B C2 C1 C2
IB + IB +
+ VCE + VCE
B
vin - vin -
RB
(a) (b)
In this circuit, the bias circuit is faulty. It is missing a resistor that supplies
a positive voltage to the base. Since there is no current through RB , we have
VB =0 and IB =0. Since VB < V0 , the BJT is in the OFF state. IC =0 and
VCE =VCC . In this case, the small-signal gain is zero.
4.5. SMALL-SIGNAL BJT MODEL 166
Example 12
Find the gain of the circuit given in Fig. 4.32(b) with RB =10 K, RC =1 K,
V0 =0.7 V, VSat =0.2 V, β=150, and VCC =12 V.
We find the DC base current, IB as
12 − 0.7
IB = = 1.13 mA (4.45)
10
Assuming that the BJT is in ACT state, we have
Since VCE < VSat , the transistor is in the SAT state, and the small-signal
voltage gain is zero.
Example 13
Find the gain of the PNP BJT circuit given in Fig. 4.33(a) with the same values
of bias circuit example at page 160: RB1 =33 KΩ, RB2 =8.2 KΩ, RC =2.2 KΩ,
RE =680 Ω, VCC =15 V, V0 =0.7 V, and VSat =0.2 V while β varies in the range
90–180. C1 and C2 are large capacitors and their effect can be neglected for the
gain calculation.
VCC
IE RE
RB2
+ ib
C1 IB V- 0 - vout
VCE
rbe βib
+ + C2 +
vin RB1 vout vin RT
IC RC RC
RE
(a) (b)
Figure 4.33: (a) PNP BJT amplifier circuit, (b) Equivalent circuit for gain
calculation.
Solution
The bias circuit was analyzed earlier in page 160: For β=90, IB =0.0336 mA, and
for β=180, IB =0.0177 mA. Hence we determine the value of rbe for the range
of β: For β=90, rbe =0.77 KΩ, and for β=180, rbe =1.46 KΩ. The small-signal
equivalent circuit is shown in Fig. 4.33(b). Here, RT stands for the parallel
4.5. SMALL-SIGNAL BJT MODEL 167
combination of the bias resistors, RB1 and RB2 . Since vin =rbe ib + RE (β + 1)ib ,
we have
vin
ib = (4.48)
rbe + (β + 1)RE
For β=90, ib =vin /62.6 K, and for β=180, ib =vin /124 K. Since vout =−RC βib ,
the gain vout /vin is −3.16 for β=90, and −3.19 for β=180. The presence of the
emitter resistor, RE , reduces the gain, but it makes it almost independent of
the value of β.
4.6 Examples
Example 14
Find Vo versus Vin for the circuit shown in Fig. 4.34, assuming that the diode
is ideal.
+ 680 VD + ID +
-
+
820 5V
Vin V o
10mA
- 100 -
Solution
1. Assume that the diode is OFF. We can solve the remaining part using
nodal analysis:
Vo − Vin Vo
+ − 10 = 0 or Vo = 0.547Vin + 3.72
0.68K 0.82K
This equation is valid as long as Vo < 5 or when Vin < 2.34 V. We verify
that the diode is OFF:
VD = Vo − 5 < 0
2. For Vin ≥ 2.34 V, the diode is ON. In this case, we write the nodal equation
as
Vo − Vin Vo Vo − 5
+ − 10 + = 0 or Vo = 0.116Vin + 4.728
0.68K 0.82K 0.1K
We can verify that for Vin ≥ 2.34 V, the diode has a positive current
through it:
Vo − 5
ID = = 1.16Vin − 2.71 ≥ 0 for Vin ≥ 2.34 V
0.1K
Example 15
Find the inductance current, iL (t), and the voltage vo (t) as a function of time.
Suppose that the initial inductance current, iL (0)=0, and the diodes have a
turn-on voltage of 0.7 V.
4.6. EXAMPLES 169
D2 VZ=15 V
Z
L =3 mH
iin(t) iL
20 mA - vL +
+ +
iin(t) D1
vo(t) 12 V
10 mA
-
10 µs
Solution
We have iin (0+ ) = 20 mA. Since iL (0+) = 0, the current can only flow
through D1. D2 and Z remains OFF.
Example 16
Find the output voltage of the half-wave rectifier circuit given in Fig. 4.37 in
terms of given parameters (ω = 2π/T ). Assume that L is very large and the
diodes are ideal.
4.6. EXAMPLES 170
iin(t)
20 mA iL(t)
10 mA
t2 t3
t1 10 µs t
27.7
vo(t)
12
t1 t2 t3
-0.7 10 µs t
D1 + vL(t) - iL (t)
+ L +
Vp sin (ωt) D2 RL vo (t)
-
Solution
Since the inductance is very large, we assume that the inductance current,
iL (t) = IL , is a constant, and the output voltage is vo (t) = RL IL is also a
constant. In the steady-state, the average voltage across the inductor should be
zero. Otherwise, the inductance current may increase indefinitely. Since IL > 0,
one of the diodes should be turned on to carry that current. When the input
voltage Vp sin ωt is greater than zero, D1 turns on. When the input voltage is
negative, D2 turns on. Hence we write the inductor voltage as
(
Vp sin(ωt) − vo if Vp sin(ωt) > 0
vL (t) =
−vo if Vp sin(ωt) < 0
Let us find the average value of vL (t) by finding the integral over one complete
cycle:
!
1 T
Z Z T /2 Z T
1 2π Vp vo vo
vL (t) = (Vp sin( t) − vo )dt + (−vo )dt = − −
T 0 T 0 T T /2 π 2 2
Hence we find
Vp
vo (t) =
π
Note that the half-wave rectifier circuit of Fig. 4.9 using a capacitor as the
voltage smoothing element has a much larger output voltage of vo (t) = Vp . So,
the circuit of Fig. 4.37 is very rarely used.
4.6. EXAMPLES 171
Example 17
vC (t)
+
+ +
10 sin (ωt) C D vo (t)
-
Assuming that the initial capacitor voltage, vC (0) = 0, find the output
voltage, vo (t), in the voltage clamper circuit of Fig. 4.38. Assume that the diode
has a turn-on voltage of 0.6 V.
Solution
In the first positive half cycle of the input, the diode remains off, and the
capacitor voltage remains at zero. Hence we have vo (t) = 10 sin(ωt) during this
time. In the second half-cycle when 10 sin(ωt1 ) < −0.6, the diode turns on.
We have vo (t) = −0.6 V. The capacitor voltage charges to the peak voltage of
vc = 10 − 0.6 = 9.4 V when 10 sin(ωt2 ) = −10, just like the half-wave rectifier.
For t > t2 , we have
10 sin(ωt)
for 0 < t < t1
vo (t) = −0.6 for t1 < t < t2
10 sin(ωt) + 9.4 for t > t2
For t > t2 , we have a shifted sine wave, whose negative peaks are clamped to
−0.6 V as shown in Fig. 4.36.
19.4
vo (t)
10
t1 t2
-0.6 t
10 sin (ωt)
Figure 4.39: The input and output voltages of voltage clamper of Example 17.
Example 18
Assuming that the initial capacitor voltages, vC1 (0) = vC2 = 0, find the output
voltage, vo (t), in the voltage doubler circuit of Fig. 4.40. Assume that the diodes
have a turn-on voltage of 0.6 V.
4.6. EXAMPLES 172
+ D2
+ + +
10 sin (ωt) C1 D1 v1 (t) C2 vo (t)
- -
Solution
C1 and D1 form a voltage clamper as investigated in Fig. 4.38. D2 and C2
form a half-wave rectifier of Fig. 4.9. After one and a quarter cycles, the output
voltage becomes nearly double the peak voltage of the input voltage, as shown
in Fig. 4.41.
18.8
v1 (t)
vo (t)
9.4
t1 t2
-0.6 t
10 sin (ωt)
Figure 4.41: The input and output voltages of voltage doubler of Example 18.
Example 19
L
+ vL - D Vo
Vin
iL +
S C Io
Referring to Fig. 4.42, suppose that switch is turning on and off at a fre-
quency of 1/T . D is a Schottky diode with a forward voltage drop of VF . C
is very large. The output current is sufficiently large so that the current iL
remains positive at all times. With Vin < Vo , find the duty cycle (ton /T ) in the
steady-state.
4.6. EXAMPLES 173
Solution
Since C is very large, the output voltage Vo is assumed to be constant without
any ripple. In the first part of the cycle, when the switch is turned on, the
inductor has a constant voltage of Vin across it. Using the equation of the
inductor
diL
vL = L
dt
we find that the current of the inductance increases linearly starting from an
initial positive value of Ii :
Vin
iL (t) = t + Ii
L
At t = ton , the change in the current is
Vin
∆iL = ton
L
In the second part of the cycle, when the switch is turned off, the current
of the inductance continues to flow. During that time, the voltage across the
inductor becomes vL = Vin − (VF + Vo ) < 0. Since this is a negative and
constant value, the current of the inductance decreases linearly. The change in
the current during the time interval T − ton is given by
(Vin − VF − Vo )
∆i′L = (T − ton )
L
In the steady-state, the increase in the current during the first part of the cycle
(ton ) should be equal to the decrease in the current during the second part of
the cycle (T − ton ). We must have ∆iL = −∆i′L . Hence we have
Vin (Vin − VF − Vo )
ton = − (T − ton )
L L
or
ton VF + Vo − Vin
= (4.49)
T VF + Vo
Since the average of the inductor current should be equal to the load current,
we also have
Io = Ii + ∆iL /2
Example 20
If a circuit has a nonlinear element, like a diode, we can still use Thèvenin or
Norton equivalent circuit method for the linear parts of the circuit. In this case,
the nonlinear element must be left out of the dashed-box as demonstrated in
this example.
Consider the circuit given in Fig. 4.43(a) where v1 (t) is an arbitrary function
of time. Let us find the Thévenin equivalent of the resistive parts of the circuit:
Applying the procedure:
4.6. EXAMPLES 174
R1 A D B R1 A vth B
+ -
+ R2 + R2
v1(t) − R3 v1(t) − R3
(a) (b)
Req
R1 A B Req A B
+ -
R2 vth (t) +
R3 −
(c) (d)
Figure 4.43: Example for Thévenin equivalent circuit for a circuit containing
nonlinear element
1. Using the voltage divider formula of Eq. 2.30, we find the open-circuit
voltage at terminals A-B with diode removed (Fig. 4.43(b)):
R2
vth (t) = v1 (t)
R1 + R2
Note that R3 does not have an effect here, since there is no current through
it.
2. Kill the voltage source as in Fig. 4.43(c). Find the resistance between the
terminals A and B (also with diode removed):
R1 R2
Req = + R3
R1 + R2
Example 21
For the NPN BJT circuit shown in Fig. 4.44(a), we have R1 =22 KΩ, R2 =8.2 KΩ,
R3 =120 Ω, VCC =12 V, β=150, V0 =0.7 V for BJT and diodes, and VSat =0.2 V.
Find the value of Rc to set VCE =4 V. With this value of Rc , what is VCE if
β=300?
Solution
We assume that the BJT is in ACT state since VCE = 4 > VSat =0.2 V. We find
the Thevenin equivalent circuit (Fig. 4.44(b)) with
R2 R1 R2
.VT = VCC = 3.26 V and RT = = 5.97 KΩ
R1 + R2 R1 + R2
From KVL at the base side going to ground via emitter path:
VT = IB RT + V0 + 2V0 + R3 IE
4.6. EXAMPLES 175
VCC VCC
R1 RC IC RC
+ +
VCE VCE
IB IB
- RT -
D1 D1
R2 VT
D2 D2
IE
R3 R3
(a) (b)
Therefore, Rc =794 Ω.
If β is changed to 300, we assume BJT is ACT and we get
3.26 − 2.1
IB = = 0.0276 mA
5.97 + (300 + 1)0.12
With IC = βIB = 8.26 mA and IE = (β + 1)IB = 8.29 mA, we get
Example 22
Find and plot vC (t) for the circuit shown in Fig. 4.45 in the interval 0 < t < 2 ms.
We have VCC =24 V, vC (0) = 0 V, C=150 nF, R1 =82 KΩ, R2 =390 Ω, β=220,
V0 =0.7 V, Vsat =0.2 V, and vIN =3.3 V for t > 0.
Solution
Since vIN > 0.7 V, the BJT is not OFF. We write for the base circuit:
vIN = R1 IB + V0 + R2 IE = R1 IB + V0 + R2 (β + 1)IB
4.6. EXAMPLES 176
VCC
+
C vC
-
R1
vIN +
−
R2
Hence
vIN − V0 3.3 − 0.7
IB = = = 0.0155 mA
R1 + R2 (β + 1) 82 + 0.39 × 221
Now, assume that the BJT is ACT at t = 0. Hence IC = βIB = 3.40 mA and
IE = (β + 1)IB = 3.42 mA. At t = 0 we have VCE = VCC − vC (0) − R2 IE =
24 − 0 − 0.39 ∗ 3.42 = 22.7 V. Since VCE = 22.7 > Vsat = 0.2 V, the BJT is ACT
at t = 0. The capacitor is charged with a constant current of IC = 3.40 mA.
Since
dvC
IC = C
dt
integrating both sides, we have
IC 3.40 × 10−3
vC (t) = vC (0) + t=0+ t = 2.27 × 104 t
C 150 × 10−9
The collector to emitter voltage of BJT is given by
VCE (t) = VCC − vC (t) − R2 IE = 22.7 − 2.27 × 104 t
So, VCE decreases linearly with time. The BJT becomes SAT when VCE =
Vsat = 0.2 V at some t1 > 0:
VCE (t1 ) = 22.7 − 2.27 × 104 t1 = 0.2
or t1 = 1 ms. What happens after t > t1 ? Consider the equivalent circuit
shown in Fig. 4.46. We have a first order circuit with an exponential solution
with vC (t1 ) = 22.7. We find
R2 0.39
vC (∞) = VCC − Vsat − vIN = 24 − 0.2 − 3.3 = 23.8
R1 + R2 82 + 0.39
and the time constant is τ = C(R1 ∥ R2 ) = 58.2 µs. Therefore
vC (t) = 23.8 + (22.7 − 23.8)e−(t−t1 )/58.2µ for t > t1
vC (t) is plotted in Fig. 4.47 for both the ACT (the voltage increases linearly)
and the SAT regions of BJT (the voltage increases exponentially).
4.6. EXAMPLES 177
VCC
+
C vC
-
+
R1 Vsat
vIN +
−
R2
Figure 4.46: Circuit for Example 22 after the BJT became SAT.
vC
23.8
22.7
ACT SAT
0 t1 t
Example 23
For the NPN BJT circuit shown in Fig. 4.48(a), find the small-signal gain
Av = vc /ei n. We have VBB =3.3 V, R1 =47 KΩ, R2 =1 KΩ, V0 =0.7 V, β=85,
VCC =15 V, VSat =0.2 V.
Solution
We first make a DC analysis by assuming that the BJT is ACT.
VBB − V0 3.3 − 0.7
IB = = = 0.055 mA
R1 47
Hence, IC =4.7 mA. We find VCE = VCC − IC R2 =15−4.7×1=10.3 V. Since
10.3 > 0.2, ACT assumption is valid.
We draw the small-signal equivalent circuit as in Fig. 4.48(b) by killing the
DC sources, VBB and VCC , and with
0.0259
rbe = = 0.468 KΩ
0.055 mA
We find the small-signal base current, ib , as
ein ein
ib = =
R1 + rbe 47.47
4.6. EXAMPLES 178
VCC
R2
R1 + R1
vC =VC+vc vc
+ +
ein iB=IB+ib - ein ib rbe βib R2
+
VBB
(a) (b)
4.7 Problems
1. For the circuit shown in Fig. 4.49, the initial voltage of the capacitor is
vC (0)=0 V, and the threshold voltage of the diode, Vo =0.7 V. Find vC (t)
for t > 0 and plot it. Make sure that you specify the value of vC (2 ms)
(It is the variable that stays continuous at discontinuities.)
1K Vin
3V
+ +v
Vin − 2µF C
2ms t
2. For the circuit of Fig. 4.50, the switch S is closed for 0 < t < 5 µs, and
then it is kept open. The initial value of the inductor current is iL (0) = 0.
Find and plot the current of the inductor, iL (t), as a function of time
between 0 < t < 15 µs. Assume that the diode has a forward voltage drop
of 0.7 V, when it is ON.
100µH
S iL
+ +
12 V 5V
D
C1 + C3 +
+ +
10 sin (ωt) D1 D2 v1 (t) D3
-
C2 + C4 +
- vo (t) +
5. Determine the states of BJT and collector to emitter voltages, VCE in the
circuit of Fig. 4.28 at p. 161 with resistor values RB1 =22 KΩ, RB2 =3.3 KΩ,
RC =1.5 KΩ, and with the parameters V0 =0.7 V, VSat =0.2 V, VCC =12 V,
and
(a) β=100
(b) β=200
(c) β=400
6. Find the state of the BJT and the voltage, VC , in the circuit of Fig. 4.52
with resistor values RB1 =2.2 KΩ, RB2 =27 KΩ, RE =100 Ω, RC =1.5 KΩ,
and with the parameters V0 =0.7 V, VSat =0.2 V, VCC =12 V, and
VCC
RE
RB1
+
RB2 RC VC
-
(a) β=70
(b) β=150
(c) β=250
7. Determine the state of the BJT and the collector to emitter voltage, VCE in
the circuit of Fig. 4.53(a) with resistor values RB1 =15 KΩ, RB2 =3.9 KΩ,
RB3 =56 KΩ, RC =820 Ω, RE =390 Ω and with the parameters V0 =0.7 V,
VSat =0.2 V, VCC =12 V and β=220.
8. Determine the collector voltage, VC in the circuit of Fig. 4.53(b) with resis-
tor values RB =120 KΩ, RC =1.5 KΩ, RE =220 Ω and with the parameters
V0 =0.7 V, VSat =0.2 V, VCC =6 V, VEE =−6 V, and β=300.
4.7. PROBLEMS 181
VCC VCC
IC IC
RB1 RC RC
IB + IB +
VC
+ VCE + -
RB3 V0 - RB V0
- -
RB2 IE RE IE RE
VEE
(a) (b)
9. Determine the state and the collector to emitter voltage, VCE in the circuit
of Fig. 4.54 with resistor values RB1 =100 KΩ, RB2 =8.2 KΩ, RC1 =3.9 KΩ,
RC2 =2.2 KΩ, and with the parameters V0 =0.7 V, VSat =0.2 V, VCC =15 V
and β=200.
VCC
RC1
RB1 A
RC2
IC
IB +
B VCE
+
RB2 V0 -
-
10. Find the small-signal voltage gain, AV , of the circuit of Fig. 4.30 at p. 163
with RB =560 K, RC =820 Ω, V0 =0.7 V, VSat =0.2 V, VCC =15 V and
β=150. First, find the base current and the state of the transistor.
11. Find the small-signal voltage gain, AV , of the circuit of Fig. 4.31 at
p. 165 with RB1 =27 K, RB2 =3.9 K, RC =680 Ω, V0 =0.7 V, VSat =0.2 V,
VCC =15 V and β=150. First, find the base current and the state of the
transistor.
12. We have a NPN BJT with parameters V0 =0.7 V, VSat =0.2 V while β=180–
460. Design a BJT amplifier with a small-signal voltage gain of −100±5
using a supply voltage of 15 V.
4.7. PROBLEMS 182
13. Find the small-signal voltage gain of the circuit of Fig. 4.55(a) with IB =40 µA,
β=150, RC =1 K, RE =220 Ω. Assume that the transistor is in ACT state.
VCC VCC
IC
RB1 RC RB1
IB vout IB
+ C + C
2 2
vin + VCE vin + VCE
C1 V0 - C1 V0 -
- - vout
RB2 IE RE RB2 IE RE
(a) (b)
14. Find the small-signal voltage gain of the circuit of Fig. 4.55(b) with
IB =40 µA, β=150, RE =560 Ω. Note that BJT in this circuit cannot
be in SAT state.
Chapter 5
TUNED CIRCUITS
1
fo = √ (5.3)
2π LC
25330
fo2 (MHz2 ) = (5.4)
L(µH)C(pF)
flywheel
water
water tank
shown as the diaphragm pushes the water. The flywheel gains speed, and it takes
over when the diaphragm is released. The flywheel pushes the water, and the
diaphragm is stretched in the other direction. When the flywheel stops, the
diaphragm pushes the water in the other direction. Hence, the back-and-forth
movement of the water continues at the system’s resonance frequency. The
resonance frequency is determined by the moment of inertia of the flywheel and
the stiffness of the diaphragm.
LC circuit of Fig. 5.1(a) does not contain any loss. To introduce a loss, we
should add a resistance to the circuit. Consider a slightly modified circuit given
in Fig. 5.1(b). The impedance now becomes
−1
1 1 jωLR jωL
Zp (ω) = + jωC + = =
R jωL R + jωL − ω 2 RLC jωL/R + (1 − ω 2 LC)
(5.5)
This
√ impedance is complex, and it indicates a resonance. At frequency ω o =
1/ LC, Z(ω) becomes
jωo L
Zp (ωo ) = =R (5.6)
jωo L/R + 0
5.1. PARALLEL RLC CIRCUIT 185
Vp = Ip R (5.7)
where Vp is the output voltage phasor. In the time domain, this is Vp cos ωo t =
Ip R cos ωo t. This means that the current through the resistor is exactly equal
to the input current at resonance frequency ωo .
At the resonance, the current phasor in the capacitor branch, IC is given by
Vp Ip R
IC = = = Ip (jωo RC) (5.8)
1/jωo C 1/jωo C
If the factor ωo RC is greater than one, we have |IC | > Ip . The capacitor
current can be larger than the current supplied by the source! In the time
domain representation, we find
IC
Imag Imag IC
IC
IL +IC
Ip =IR IR
Ip
Real Real Ip Real
IR
IL +IC
IL
ω=ωo ω=ω1 IL ω=ω2
IL (a) (b) (c)
At the resonance, the inductor current has the same magnitude as the capac-
itor current, but it is 180o out of phase. That is why the parallel combination
of LC acts like an open-circuit: The current through the inductor cancels the
current through the capacitor.
The resonance is an interesting phenomenon!
5.1. PARALLEL RLC CIRCUIT 186
where the approximation is good within 1%. The phasor diagrams showing
the current phasors at these frequencies are given in Fig. 5.3(b) and (c). The
difference between these frequencies is
fo
∆f = f2 − f1 = (5.25)
Q
is called the 3-dB bandwidth (BW) of the tuned circuit. The BW of the tuned
circuit in the figure is 1/10 of resonance frequency since Q is chosen √
as 10. From
Eq. 5.23, we can see that the angular resonance frequency, ωo = 1/ LC, is the
geometrical mean of ω1 and ω2 :
√
ωo = ω1 ω2 or fo2 = f1 f2 (5.26)
Figure 5.5: |Zp (ω)| of two parallel tuned circuits with Q=1 and Q=10.
Q=10, as in Fig. 5.4, is plotted together with the impedance of a circuit with
Q=1, for comparison. Both circuits have the same C and L values, hence the
same ωo , but the parallel resistance of the
√ high-Q circuit is 10 times larger than
the other one. Note that |Zp (ω)| is R/ 2 = 0.7R at ω = ω1 and ω2 .
The ratio of the voltage magnitudes generated across the circuit in Fig. 5.1(b)
at resonance and ω1 (or ω2 ) is
|V (ωo )| |Zp (ωo )| √
= = 2 (5.27)
|V (ω1 )| |Zp (ω1 )|
and
V (ωo ) √
20 log10 = 20 log10 2 = 10 log10 2 = 3dB (5.28)
V (ω1 )
This is why the bandwidth ∆f = f2 − f1 is called 3-dB BW.
5.2. SERIES RLC CIRCUIT 189
Example 1
Consider RLC circuit as in Fig. 5.1(b) with C = 100 pF and R=500 Ω. Find
the value of L to resonate the circuit at 28 MHz. Find the quality factor, Q, and
determine IC , IL , and IR at resonance if Ip =1 mA. Find the 3 dB frequencies.
From Eq. 5.4 we find
25330 25330
L(µH) = = = 0.323 µH
fo2 (MHz2 )C(pF) 282 · 100
IL = −IC = −j8.80 mA
and
s
2
1018 1012 1012
1
f2 = + + = 29.64 MHz
2π 0.323 · 100 2 · 500 · 100 2 · 500 · 100
As a check we use Eq. 5.26: 26.45 · 29.64 = 784.0 = 282 . On the other hand,
simpler expressions of Eq. 5.24 give good approximations:
1 1
f1 ≈ 28 1 − = 26.41 MHz and f2 ≈ 28 1 + = 29.59 MHz
2 · 8.8 2 · 8.8
1 1 − ω 2 LC
Zs (ω) = + jωL = (5.29)
jωC jωC
5.2. SERIES RLC CIRCUIT 190
√
Zs (ω) is zero when 1 − ω 2 LC=0 or ω = ωo = 1/ LC. This frequency is the
series resonance angular frequency.
When we add a series loss element R into the circuit, we obtain the circuit
in Fig. 5.6(b). The impedance of the series RLC circuit is
1 1 − ω 2 LC + jωRC
Zs (ω) = + jωL + R = (5.30)
jωC jωC
At the angular resonance frequency, ωo , we have Zs (ωo ) = R. At this angular
frequency, the effect of series L and C cancels, hence only R remains. If we
apply a voltage of Vs cos ωo t across the RLC circuit, the current is Is cos ωo t =
(Vs /R) cos ωo t. Hence, in phasor notation, we write
Vs
Is = at ω = ωo (5.31)
R
The voltage phasors across the inductor and capacitor are given by
jωo L Is jωo L
VL = Is jωo L = Vs and VC = = −Is jωo L = −Vs = −VL
R jωo C R
(5.32)
Therefore, the voltage across the inductor has the same magnitude but opposite
polarity of the voltage across the capacitor. The voltages add up to zero at
resonance. Also, notice that the magnitude of the inductor or capacitor voltage
can be greater than the applied voltage Vs , if the factor jωo L/R is greater than
one.
3-dB angular frequencies are given by
s 2 s 2
1 R R 1 R R
ω1 = + − and ω2 = + + (5.33)
LC 2L 2L LC 2L 2L
Es 2πfo L ωo L 1
Q = 2π = = = (5.40)
Ed R R ωo RC
Note that the Q factor for the parallel circuit given in Eq. 5.21 is the inverse
of this factor. To have a large Q factor, we need to increase L to increase the
stored energy, and reduce R to reduce the dissipated energy.
The impedance of a series RLC circuit can be written as
Figure 5.7: Real and imaginary parts of Zs (ω) normalized to series resistor R.
Figure 5.8: Real and imaginary parts of Zs (ω) normalized to series resistance
R.
resistor. This kind of circuit is called the tank circuit. As far as the resonance
is concerned, this circuit can be viewed as a series resonance circuit containing
series-connected C, LS , and RS . However, we are interested in what appears
across the capacitor terminals. The admittance of the tank circuit is
1 RS RS2 C − LS + ω 2 L2S C
YT (ω) = jωC + = 2 + jω (5.42)
RS + jωLS RS + ω 2 L2S RS2 + ω 2 L2S
At the resonance, the imaginary part must be zero:
RS2 C − LS + ωo2 L2S C = 0 (5.43)
This condition yields
" 2 #1/2 1/2
1 RS 1 1
ωo = − = (5.44)
LS C LS LS C 1 + 1/Q2
Equating the real part of the left-hand-side to the real part of the right-hand-
side, we find
Rp (ωo Lp )2 Rp
RS = 2 = 2 (5.46)
Rp + (ωo Lp )2 Rp /(ωo Lp )2 + 1
Since Rp /ωo Lp = Q, we have
Rp
RS = or Rp = RS (Q2 + 1) (5.47)
Q2 + 1
Rp2 ωo Lp ωo Lp
ωo LS = = (5.48)
Rp2 + (ωo Lp )2 1 + (ωo Lp )2 /Rp2
This equivalence also maintains that the Q’s of two circuits are also equal:
Rp ωo LS
Q= = (5.50)
ωo Lp RS
Example 2
We have an inductor with L=0.300 µH and a series resistor of 1 Ω. Design a
resonant circuit using this inductor at 8.00 MHz. Find an equivalent parallel
RLC circuit.
From Eq. 5.50, with LS =0.300 µH and RS =1 Ω, we find at f =8.00 MHz
Q=15.1. From Eq. 5.49, we get with LS =0.300 µH, Lp =0.301 µH. Hence we
must have C=1314 pF. The parallel resistance, Rp is found from Eq. 5.47 as
Rp =229 Ω.
Es (1/2)CVp2
Q(f ) = 2π = 2π = 2πf RC = ωRC (5.52)
Ed To (1/2)Vp2 /R
5.4. REAL INDUCTORS 194
The equation above suggests that the Q factor increases as the frequency in-
creases. However, the insulator resistance decreases as the frequency is in-
creased. As a result, capacitors also have an optimum frequency where the
quality factor is maximized.
Figure 5.9: (a) A capacitor, (b) the capacitor model with parasitic inductance,
(c) the full high frequency model
given in Fig. 5.9(c). The parallel resistance Rp models the loss in the dielectric
material from which the capacitor is made up. The series resistor Rs represents
the sum of conductor resistance in the leads and the losses at the lead contacts.
Usually, Rp is very high and can be ignored.
d2 N 2
L(µH) = (5.60)
46 d + 102 l
where L is the inductance value in µH, d is the coil diameter in cm, N is the
number of turns, and l is the length of the coil in cm. This formula is accurate
for coils having an aspect ratio l/d greater than 0.4.
Example 3
Let us find the number of turns and the dimensions for a 0.33 µH inductor.
There are many choices for dimensions. Let us choose a coil geometry such that
the length of the coil is equal to its diameter, i.e., l = d. In this case, we have
dN 2
L(µH) = if l = d (5.61)
148
Once it is wound, the value of an air-core inductor can be tuned within about
20%, by extending its length. As the length is increased, the inductance value
reduces.
l
Rdc = ρ (5.62)
A
where ρ is the resistivity of the wire material (refer to Table 2.1 on page 19), l is
the length, and A is the cross-section of the wire. This loss is further aggravated
at RF because of a phenomenon called skin effect. As the frequency increases,
the current is no longer homogeneously distributed across the cross-section of
the conductor. It is confined to a thin cylindrical layer next to the conductor
5.4. REAL INDUCTORS 198
When the wires are next to each other, there is an additional effect called
proximity effect which causes the current distribution in the wires be affected
by the current in the neighboring wires increasing the resistance even further.
Fig. 5.15 shows the current distribution in three neighboring 0.35 mm diameter
wires at 3 MHz.
5.4. REAL INDUCTORS 199
Example 4
Let us determine the resistance of 10 cm of 1 mm diameter copper wire at
28 MHz. From 19, copper has a resistivity of ρ = 1.68 · 10−8 Ωm. Hence, a
10.0 cm copper wire of 1.00 mm diameter has a DC resistance of 2.14 mΩ. fsc
for this wire is found from Eq. 5.63 as 0.08 MHz, or 80 kHz. Then, at 27 MHz
R(f ) becomes 39.3 mΩ.
Es (1/2)LIp2 2πf L ωL
Q(f ) = 2π = 2π 2
= = (5.66)
Ed To (1/2)RIp R R
Eq. 5.66 suggests that the quality factor improves as the frequency is increased.
This is true if the resistance R remains constant as frequency is changed. How-
ever, the equivalent loss resistor of the inductor increases as the frequency is
5.4. REAL INDUCTORS 200
Figure 5.16: (a) An inductor, (b) the inductor model with parasitic capacitance
Example 5
Consider an inductor made by winding a single layer of 32 turns on a core
with an AL of 20 nH/turn2 . Hence the inductance is 20.5 µH. The inductor’s
5.4. REAL INDUCTORS 201
5.4.7 RF choke
A choke is an inductor used to block higher-frequency current while letting low
frequency or DC current pass. If the intended blocking frequencies are RF, the
inductor is called RF choke or RFC. While the core loss degrades the perfor-
mance of an inductor used in resonant circuits, it provides useful properties in
a choke. A particularly important application area is radio frequency interfer-
ence (RFI) or electromagnetic interference, EMI (EMI). The interference of RF
signals within the same instrument or between instruments must be avoided.
Voltages or currents can couple by electromagnetic means to other parts of the
circuits where they are not wanted. Interference of an irrelevant signal in a
circuit can cause an instrument to malfunction.
A typical RF interference to the power supply line is shown in Fig. 5.17(a).
Here, a circuit is fed by a DC power supply. There is a bypass capacitor,
Figure 5.17: (a) An unplanned interference signal Vint being applied to the low
frequency circuits, (b) adding a series R to form a low-pass-circuit with some
loss at DC, (c) adding an RFC to form a low-pass-filter with no loss at DC.
by the circuit. This is not always agreeable, particularly in circuits where the
DC current demand is high or varies considerably.
A better solution is to use an inductor instead of a resistor, as shown in
Fig. 5.17(c). All we need in this inductor is that it must exhibit a high impedance
at the frequency of Vint , and a low impedance (preferably zero impedance) at
DC. Therefore, it need not be a high Q inductor. Such an inductor is an RFC.
A common way of making RFC is to wind a few turns on a ferrite core
with high permeability. The inductor behaves like a simple inductance at low
frequencies, and its impedance is zero at DC, since the core loss of the ferrite at
low frequencies is zero. The DC supply voltage, Vdc , appears at the terminals
of the LF circuit.
The impedance of the RFC at HF, on the other hand, is obviously
where Rc is the resistance due to core loss. Since the existence of Rc makes
|Zω| larger, having a large core loss is preferable for this application. The effect
of Vint at circuit terminals is what is left after the voltage division between Z
and 1/(jωC):
Vint
(5.69)
1 − ω 2 LC + jωRc C
A modern way to eliminate RF noise is to use a ferrite bead, a cylindrical-
shaped core of ferrite slipped over a wire. A typical ferrite material made from
nickel-zinc alloy used for this purpose might have µr =1000, and has a large loss
factor above 1 MHz. Computer power cords often have such chokes, consisting
of cylindrical ferrites encircling the cords to block noise.
5.5 Transformers
Transformers are the most common impedance transformation devices. The
ideal transformer of Section 2.12 is depicted in Fig. 5.18 again, using phasor
notation. The primary and secondary voltage and current relations in an ideal
I1 n1 : n2 I2
+ +
Zin V1 V2 ZL
- -
transformer are
V2 n2 I2 n1
= and = (5.70)
V1 n1 I1 n2
where n2 /n1 is the ratio of the number of turns in the secondary winding to the
number of turns in the primary, commonly called the transformer turns ratio.
5.5. TRANSFORMERS 203
The impedance ZL gets transformed by the (n1 /n2 )2 ratio. Indeed, any shunt
(i.e., parallel) or series impedance on the secondary can be transferred to pri-
mary side as a shunt or series element provided their value multiplied by (n1 /n2 )2 .
Similarly, any impedance on the primary side can be transferred to the secondary
side through multiplication by the factor (n2 /n1 )2 .
Example 6
Consider the ideal transformer shown in Fig. 5.19(a). The series impedance Z1
on the primary can be moved to the secondary upon multiplication by (n2 /n1 )2
(see Fig. 5.19(b)). Similarly, the shunt impedance Z2 is moved to the secondary
upon multiplication by the same factor as depicted in Fig. 5.19(c). All the
circuits shown in the figure are equivalent to each other as far as other circuits
are concerned.
+ Z1 + + Z1(n2/n1)2 + + Z1(n2/n1)2 +
V1 Z2 V2 V1 Z2 V2 V1 Z2 (n2/n1)2 V2
n1 : n2 n1 : n2 n1 : n2
(a) (b) (c)
currents is the difference between primary and secondary. The flux, Φ, generated
in the core by two currents Ip and Is is
Φ = np AL Ip − ns AL Is (5.71)
where np and ns are the number of turns in primary and secondary windings,
respectively, AL is the inductance constant of the core. We can write Ip as
ns Φ
Ip = Is + (5.72)
np n p AL
The first term is the relation between the primary and secondary currents in
an ideal transformer. It is called the transformer current. The second term is
called the magnetizing current and it is related to the finite value of the primary
inductance (note that this inductance is implicitly assumed to be infinite in the
ideal transformer).
The same flux induces voltages across the primary and secondary windings.
The induced primary and secondary voltages, on the other hand, are
Vp = jωnp Φ and Vs = jωns Φ (5.73)
Using these relations, we can relate two voltages as
Vp np
= (5.74)
Vs ns
Substituting Vp = jωnp Φ into Eq. 5.72 we obtain
ns Vp ns Vp
Ip = Is + 2
= Is + (5.75)
np jωnp AL np jωLp
where we defined Lp = n2p AL as the inductance of the primary winding. The two
equations relating the primary and secondary terminal voltages and currents can
be shown in the form of an equivalent circuit comprising an ideal transformer and
a parallel magnetizing inductance, Lp , as in Fig. 5.21(a). When the secondary
is open-circuited (Is = 0), the equivalent circuit should be the same as the
inductor of the primary, Lp . Since the primary of the ideal transformer does
not carry any current, and the ideal transformer can be removed altogether
leaving behind just the inductance of the primary.
Equivalently, Lp can be transferred to the secondary side upon multiplication
by (ns /np )2 giving the value (ns /np )2 Lp = n2s AL = Ls which is the magnetizing
inductance of the secondary (see Fig. 5.21(b)). We should place either the
primary or the secondary inductance in the equivalent circuit, but not both.
5.5. TRANSFORMERS 205
+ Ip Is + + Ip Is +
Vp Lp Vs Vp Ls Vs
- - - -
n1 : n2 n1 : n2
(a) (b)
L1 ideal transf. L2
+ Ip Is +
Vp Lm Vs
- -
n1 : n2
(a)
Figure 5.22: Equivalent circuit of a lossless real transformer with 0 < k < 1.
Example 7
We have a core with AL =10 nH/T2 . A transformer is wound with n1 =10 and
n2 =4. Find the equivalent circuit of the transformer if the coupling coefficient
between two coils is k = 0.6.
5.6. TUNED AMPLIFIERS 206
Solution
We have Lp = 102 × 10 = 1000 nH, and Ls = 42 × 10 = 160 nH. Hence
L1 = (1 − 0.6) × 1000 = 400 nH, L2 = (1 − 0.6) × 160 = 64 nH, and Lm =
0.6 × 1000 = 600 nH.
Fig. 5.23 depicts photos of different size transformers suitable for different
frequency bands.
VCC
RB1 R C L
vout
Cc1 Cc2
vin
RB2 RE CE
ib
vin vout
R C L
rbe
RB1 RB2 βib
β jωL
Av = − 2
(5.76)
rbe (1 − ω LC) + jωL/R
β
Av (ωo ) = − R (5.77)
rbe
The small-signal gain of a tuned amplifier with a center frequency of 27 MHz is
plotted in Fig. 5.26 for IB =0.02 mA (rbe =1280 Ω), R=300 Ω, β=65, C=100 pF,
L=0.347 µH.
24
22
20
18
Gain (dB)
16
14
12
10
6
15 20 25 30 35 40
Frequency (MHz)
in the datasheet of LM358, we observe that the gain is 110 dB and constant up
to about only 20Hz. Above this frequency, the gain falls at 20 dB/decade as
the frequency increases.
where G is the closed-loop gain and BW is the 3-dB bandwidth of the amplifier.
We should inquire about the gain-bandwidth product (or unity-gain band-
width) of an OPAMP to find out suitability at high-frequency amplification.
For example, if an OPAMP has a gain-bandwidth product of 4 MHz, it is not
suitable for amplifying signals at high frequencies like 27 MHz.
We can calculate the gain more accurately at any frequency using Eq. 5.78.
For a non-inverting OPAMP amplifier, as in Fig. 3.26(b) (on page 105), we can
write the gain as
Vo 1
= (5.81)
Vin R1 /(R1 + R2 ) + jf /fT
|IL |2 RL 1 |Vin |2 RL
PL = = (5.82)
2 2 (RS + RL )2
With a given source of value Vin and source impedance RS , the power PL can
be maximized by choosing a suitable RL value. We find this optimum value by
5.8. MAXIMUM POWER TRANSFER 209
+ RS IL + ZS IL +
Vin RL Vin VL ZL
-
(a) (b)
PL will be maximized if
ZL = ZS∗ implying XL = −XS and RL = RS (5.89)
and the maximum power delivered to load impedance is
|Vin |2 |Vin |2
PLmax = = (5.90)
8RS 8RL
Since ZL = ZS∗ , this condition is called conjugate matching. The condition
means that the load reactance is chosen to resonate with the source reactance
at the operating frequency to maximize the current and that the load resistance
is equal to the source resistance to maximize the power transfer.
Under the conjugate matching condition, the available power from the source,
PA , is equal to the power delivered, PL , to the load. Therefore, the available
power is given by
|Vin |2
PA = (5.91)
8RS
5.8. MAXIMUM POWER TRANSFER 210
Example 8
Suppose we have a 1000 W audio amplifier to be used in a heavy-metal concert.
The output impedance of the audio amplifier, RS is 8 Ω. But the speaker has
a load impedance of RL =4 Ω. Find out how much power we lose by not using
a speaker of RL =8 Ω impedance.
Power ratings of audio
√ amplifiers are
√ specified, assuming RL =RS . From
Eq. 5.85, we find Vin = 8PLmax RS = 8 · 1000 · 8=253 V (peak voltage of a
sinusoid). Since RL =4 Ω, from Eq. 5.82, we write
1 |Vin |2 RL 1 2532 4
PL = = = 889 W
2 (RS + RL )2 2 (8 + 4)2
5.9 Examples
Example 9
Consider the circuit given in Fig. 5.28. What should RL value be for the maxi-
mum power dissipation in RL ? What is the amount of power dissipation?
75 9V A Req A
+ v
150 150 RL − th RL
80mA
B B
75
(a) (b)
Figure 5.28: (a) Circuit for Example 9, (b) Thévenin equivalent circuit.
Solution
First, let us find the Thévenin equivalent circuit between pins A and B. The
equivalent resistance Req can be found by killing the sources (open-circuit the
current source and short-circuit the voltage source):
Using Eq. 5.84, we must have RL = Req = 50 Ω for the maximum power
transfer. To find the power transferred, we need to find the Thévenin voltage,
Vth . Since there are two sources, we can use superposition to find vth , while RL
is disconnected.
After killing the current source, the open-circuit voltage, vAB , is found by
the voltage divider:
150 150
vAB = 9= 9 = 6.0 V
150 + (150 ∥ 150) 150 + 75
Killing the voltage source, vAB is found using the current divider formula of
Eq. 2.35 (see page 35):
75 75 0.080
vAB = 0.080 · 75 = 0.080 · 75 = 75 = 2.0 V
75 + (150 ∥ 150 + 75) 75 + 150 3
and hence vth = 6.0 + 2.0 = 8.0 V. The power, PL , dissipated on the resistance
RL = 50 Ω is
v 2 1 2
th 8 1 16
PL = = = = 0.32 W
2 RL 2 50 50
Note that there is an additional factor of 1/2 in Eq. 5.85, since the voltage
source is expressed as a phasor.
5.9. EXAMPLES 212
Example 10
Given a sinusoidal source with ω = 1.25 · 108 shown in Fig. 5.29(a) in dashed
lines. Find the values of n and Cp so that the maximum power is transferred to
the load resistor, RL . The available core has AL =10 nH/T2 . Find the maximum
power transferred to the load resistor RL .
n : 4
+ 240 +j180 RL
Lp
3 cos(ωt) 15Ω
Cp 250pF
(a)
+ 240 +j180
Lp 15(n2/16)
3 cos(ωt) 2
Cp 250(16/n )
(b)
Figure 5.29: (a) Circuit for Example 10, (b) the components on the secondary
side transferred to the primary.
Solution
We can transfer the 250 pF capacitor and the 15 Ω resistance to the primary
side as shown in Fig. 5.29(b). The resistance is multiplied by (n/4)2 while the
capacitance is divided by (n/4)2 . Hence the total shunt resistance of the load is
RT = 15(n2 /16), while the total shunt reactance is the parallel combination of
Cp + 250 · 16/n2 (in pF) with the inductance Lp . The load seen by the source
should be 240−j180 Ω for a maximum power transfer. Hence the load should
be a capacitance, CT , in parallel with the resistor, RT . We write the parallel
combination as
1 RT · 1/(jωCT ) RT RT − jωRT2 CT
RT ∥ = = =
jωCT RT + 1/(jωCT ) 1 + jωRT CT 1 + ω 2 RT2 CT2
For a conjugate match we must have
RT −ωRT2 CT
= 240 and = −180
1 + ω 2 RT2 CT2 1 + ω 2 RT2 CT2
Solving these two equations simultaneously, we find RT = 375 Ω and CT =
16 pF. Therefore,
2
n
RT = 15 = 375 or n = 20
16
Hence we have Lp = 10 nH/T2 · 202 = 4 µH. This inductance eliminates a shunt
capacitance of value
1
Cr = 2 = 16 pF
ω Lp
5.9. EXAMPLES 213
Hence
16
CT = Cp + 250 − Cr = Cp + 10 − 16 = Cp − 6 pF
n2
Since CT = 16 pF, we find Cp = 22pF. The power dissipated on RL is given by
Eq. 5.85
|Vin |2 |3|2
PL = = = 4.7 mW
8RS 8 · 240
Example 11
Spark plugs used in the ignition system of gasoline motors require 12,000 to
50,000 V to fire. Since most autos have a DC voltage of 12 V, a circuit is
necessary to generate the high voltages. Refer to the circuit given in Fig. 5.30(a),
where the switch S is normally closed, and it is opened at the time when the
ignition is desired. With Lp =10 mH and R = 6 Ω, explain how the high voltages
are obtained.
ip Ideal transformer is =-10mA
R 1 : 200 R ip
Lp is =0 Insulator Lp
50,000V
+ +
ip Spark 12V ip
12V plug
S S +
Gap Spark!
(a) (b)
Figure 5.30: (a) Ignition circuit with S closed, (b) ignition circuit right after S
is opened.
Solution
While S is closed, the time constant in the primary side is determined by τ =
Lp /R = 1.7 ms. While S is closed for a sufficiently long time (at least 5τ =
8.3 ms), a current of
12
ip = =2A
6Ω
flows in Lp . Since the spark plug is nonconducting, is = 0, and hence there is
no current in the primary of the ideal transformer. When S is opened (refer to
Fig. 5.30(b)), the continuity of inductance current dictates that the same current
must flow in the primary of the ideal transformer. Hence a current of is =
−ip (1/200) = −10 mA flows in the secondary, causing a spark in the spark plug.
Since this requires a voltage like 50,000 V, the primary voltage is momentarily
−50, 000/200 = −250 V. When the spark plug is ignited, the voltage at the
secondary drops and the current in the secondary decreases exponentially with
time.
5.9. EXAMPLES 214
Example 12
A tuned BJT amplifier is built using the circuit in Fig. 5.31 with β = 90,
V0 = 0.7 V, Vsat = 0.2 V, R1 = 68 K, R2 = 6.8 K, R3 = 390 Ω, VCC = 12 V,
AL = 8 nH/T2 (for the transformer core), n1 = 20, n2 = 10, R = 470 Ω. C1 and
C2 are sufficiently large capacitors so that they can be considered short-circuit
at the operating frequency. Find the DC base current and the state of the
transistor. Find the value of C such that the resonance occurs at f0 = 15 MHz.
Determine the small-signal output voltage in terms of vin at 15 MHz.
VCC
C
n1 n2
R1
vout
C1
R
+
vin R2 VE
R3 C2
Solution
Thèvenin equivalent circuit of R1 , R2 and VCC is found as RT = R1 ∥ R2 =
68 ∥ 6.8 = 6.2 K and VT = VCC R2 /(R1 + R2 ) = 1.09 V. Asuming the BJT is
ACT, we find the base current as
VT − V0
IB = = 0.0094 mA
RT + (β + 1)R3
ib
vout
+
vin R1 R2 rbe βib Lp C (n1/n2)2R
Figure 5.32: The small-signal model of tuned BJT amplifier for Example 12.
The tuned circuit has a quality factor of Q = 2πf0 C(n1 /n2 )2 R = 6.2. Hence the
3-dB frequencies are at f1 = f0 − f0 /(2Q) = 13.8MHz and f2 = f0 + f0 /(2Q) =
16.2 MHz. The BJT will be ACT as long as the peak voltage of vout is less than
VCC − VE = 11.7 V at f0 .
5.10. PROBLEMS 216
5.10 Problems
1. In a series RLC circuit, R=100 Ω, L=10 µH, and C=3 pF. Plot the
magnitude and the phase of the impedance as a function of ω for 0.6ωo <
ω0 < 1.5ωo .
2. A voltage 10∠0 is applied to the series circuit of Problem 1. Find the
voltage across each element for f = 28MHz.
3. A series circuit with R=50 Ω, C=39 pF, and variable inductor L has an
applied voltage V=10∠0 with a frequency of 16MHz. L is adjusted until
the voltage across the resistor is maximum. Find the voltage across each
element.
4. A series circuit has R= 50 Ω, L=1 µH, and a variable capacitor C. Find
the value of C for a series resonance at f =16MHz.
5. Given a series RLC circuit with R=10 Ω, L=0.5 µH, and C=220 pF,
calculate the resonant, lower and upper half-power frequencies.
6. Show that the resonant frequency ωo of an RLC series circuit is the ge-
ometric mean of ω1 and ω2 , the lower and upper half-power frequencies.
12. In the parallel circuit of Fig. 5.34(a), determine the resonant frequency if
R=0 and R=1 Ω. Compare them to resonant frequency when R=100 Ω.
13. In the parallel circuit in Fig. 5.34(b), find the resonant frequency fo .
100 1.3µH
+ R L
vin(t) C
14. Calculate the voltages across R, L, and C, of the series RLC circuit of
Fig. 5.35, if vin (t) = 5 cos(ωo t). Choose C such that the circuit resonates
at a frequency of fo =28 MHz. Write down the time waveform expressions
for these voltages.
15. Calculate the same voltages in problem 14, at the series circut’s upper
and lower 3 dB frequencies. Write down the time waveform expressions
for these voltages.
16. Calculate the DC resistance of a 0.2 mm diameter wire of length 10 cm.
Calculate its skin effect cut-off frequency. Calculate the approximate cop-
per loss resistance of an inductor made of this wire at 28 MHz.
17. What must be the length of a 900 nH air core inductor if it has 21 turns
and its diameter is 5 mm?
18. Find the number of turns of an air-core inductor with L=270 nH and
d = l =5 mm.
19. AL of the T25-10 toroidal core is given as 1.9 nH/T2 . Find the number
of turns required to make a 615 nH inductor using T25-10.
20. An inductor made by winding 7 turns on T20-7 toroid yields 0.13 µH with
a Q of 102 at 30 MHz. Find the approximate value of AL for T20-7 and
the equivalent series loss resistance of this inductor at 30 MHz.
5.10. PROBLEMS 218
21. Consider the all-pass probe compensation circuit given in Fig. 5.36. Show
that the equivalent input impedance, Zeq , becomes approximately equal
to the parallel connection of Cp and Rp , when the probe is compensated
(Rp =9RT for ×10 probe).
Figure 5.37: OPAMP circuits for (a) problem 23 and (b) problem 24.
5 · 105
Aol (ω) =
1 + jω/100
from its datasheet. What is the voltage gain of the non-inverting amplifier
at 400 kHz? What is its gain at 1 kHz?
24. The same OPAMP is used in an inverting amplifier configuration, as shown
in Fig. 5.37(b). What is the output voltage vout (t) when the input is
vin (t) = 2 cos(2π106 t + 30o ) volts? What is vout (t) when the input is
vin (t) = 2 cos(2π103 t + 30o )?
5.10. PROBLEMS 219
FILTERS
Filters are usually used to remove undesired components of a signal. For ex-
ample, an antenna delivers a complex signal containing many components at
a large band of frequencies. It is preferable to remove the unnecessary signal
components at frequencies other than the band of interest before the signal is
amplified.
Parallel and series tuned circuits can be used to filter such signals. The
filtering performance of such circuits is only determined by the quality factor
of the circuit. It is often necessary to have filters with improved performance
compared to what simple tuned circuits can offer. In what follows, we describe
more complex circuits with higher filtering performance.
6.1 Motivation
Any electronic filter can be visualized as a block between a source (input) and
a load (output). This is depicted in Fig. 6.1. What is expected from this is to
maintain the signal components at wanted frequencies and eliminate the ones
at unwanted frequencies, as much as possible. For this purpose, we may need
PL
GT = (6.1)
PA
where PL is the power delivered to load, and PA is the available power (see
Eq. 5.91 on page 209) from the source. In the absence of a filter and with
RL = RS , we have PL = PA or GT = 1. The transducer power gain, GT , is
given in terms of voltages and resistances as
2
PL |Vout |2 /2RL 4RS Vout
GT (ω) = = = (6.2)
PA |Vin |2 /8RS RL Vin
RS RS L
+ +
C RL RL
(a) (b)
RS L RS L
+ +
C RL C RL
(c) (d)
Figure 6.2: First order low-pass-filters using (a) a capacitor and (b) an induc-
tor. Second-order low-pass-filters using (c) an inductor and a capacitor, (d) a
capacitor and an inductor.
We note that the transducer power gain of the filter shown in Fig. 6.2(d),
where the filter is flipped, is the same as Eq. 6.13.
Transducer power gains of first and second-order filters are plotted in Fig. 6.3
on a semilog plot. The vertical axis is in decibels. Note the asymptotic behavior
of the filters above the cutoff frequency. The first-order filter has a slope of
−20 dB/dec, while the second-order filter has a −40 dB/dec slope. The second-
order filter is superior to the first-order filter in two respects:
1. Suppression of signal components at frequencies higher than ωc is signifi-
cantly improved,
6.2. POLYNOMIAL FILTERS 223
2. The signal components with frequencies less than ωc are better preserved,
or less attenuated.
How many filter elements must be used? Which kind of elements must be used?
Or, what must be the values of the elements? Modern filter theory addresses
these questions by a systematic filter design technique.
Third-order LPF
Fourth-order LPF
-10 Fifth-order LPF
Asymptotes
-20
Transfer function (dB)
-30
-40
-50
-60
10-1 100 101
f/f c
Figure 6.3: The transducer power gain of first and second-order LPFs as a
function of normalized frequency.
R L1 L3 L5 Ln-1
+ C2 C4 C6 Cn
R LPF
(a)
R L2 L4 L6 Ln
+ C1 C3 C5 C7 Cn+1
R LPF
(b)
R C1 C3 C5 Cn+1
+
L2 L4 L6 Ln R HPF
(c)
Figure 6.4: Ladder type low-pass-filter prototypes for (a) even, and (b) odd
number of elements. (c) High-pass-filter prototype with odd number of elements.
the HPF ladder in Fig. 6.4(c), the series and shunt elements are interchanged
compared to LPF, thus yielding exactly the opposite function.
-20
Transfer function (dB)
-30
-40
-50
-60
1 1.5 2 2.5 3 3.5 4 4.5 5
f/f c
Using this table for designing a filter is straightforward. Since the coefficients
are normalized component values, we must scale them for the given termination
resistance and cutoff frequency.
Low-pass-filter
We use the following procedure to design an nth-order Butterworth low-pass-
filter of cutoff frequency fc for load and source impedances of R:
1. Use the corresponding Butterworth table value to find the inductor value
as
bi R
Li = (6.16)
2πfc
n b1 b2 b3 b4 b5 b6 b7 b8
1 2.000
2 1.4142 1.4142
3 1.0000 2.0000 1.0000
4 0.7654 1.8478 1.8478 0.7654
5 0.6180 1.6180 2.0000 1.6180 0.6180
6 0.5176 1.4142 1.9319 1.9319 1.4142 0.5176
7 0.4450 1.2470 1.8019 2.0000 1.8019 1.2470 0.4450
8 0.3902 1.1111 1.6629 1.9616 1.9616 1.6629 1.1111 0.3902
2. Use the corresponding Butterworth table value to find the capacitor value
as
bi
Ci = (6.17)
2πRfc
Above the cutoff frequency, the signal is reduced by 20n dB/decade or 6n dB/oct.
The transducer power gain of an LPF can be drawn easily using the asymp-
totic lines shown in Fig. 6.3 or 6.5: On a semilog axis, where the frequency axis
is logarithmic, first draw a line with zero slope at 0 dB line indicating the asymp-
totic behavior in the passband. Then, draw a line with slope −20n dB/decade
(one decade is a frequency ratio of 1:10) passing through the (fc , 0 dB) point
showing the response in the frequency range above the cut-off frequency. The
actual response curve approaches asymptotes at low and high frequencies, pass-
ing through (fc , −3 dB) point. The asymptotes are good approximations for
f < 0.5fc and f > 2fc .
Example 1
Let us design a third-order Butterworth low-pass-filter for a cutoff frequency of
fc =2 MHz for source and load impedances of RS = RL =300 Ω. We can use
either L1 − C2 − L3 topology or C1 − L2 − C3 topology. The second one (shown
in Fig. 6.6(a)) is preferable since it uses only one inductor. Using the n = 3
values in Table 6.1 and R=300 Ω, we find
1.00 2.00 · 300
C1 = C3 = 6
= 265 pF and L2 = = 48.0µH
2π300 · 2 · 10 2π2 · 106
We can estimate the performance of the filter using Fig. 6.5: Since fc =2 MHz,
RS L2 RS C1 C3 C5
+ +
Vin C1 C3 RL Vin L2 L4 RL
(a) (b)
Example 2
We require a low-pass-filter for R = 50Ω, which passes the signals lower than
1 MHz with attenuation less than 1 dB and attenuates signals higher than
3.5 MHz with attenuation of more than 30 dB.
Using the decibel version of the Butterworth filter equation of Eq. 6.14 we
write the requirements as
2n
1 · 106
PL 1
10 log10 = 10 log10 ≥ −1dB ⇒ ≤ 0.259
PA 1 + (1 · 106 /fc )2n fc
6.2. POLYNOMIAL FILTERS 227
2n
3.5 · 106
PL 1
10 log10 = 10 log10 ≤ −30dB ⇒ ≥ 999
PA 1 + (3.5 · 106 /fc )2n fc
Taking the logarithm of both sides and solving two equations under equality
case, we write
1 · 106
2n ln = 27.63n − 2n ln fc = ln 0.259 = −1.35
fc
3.5 · 106
2n ln = 30.13n − 2n ln fc = ln 999 = 6.91
fc
Subtracting two equations from each other, we find n=3.29. Using one of the
equations, we reach fc =1.22 MHz. For filter order, we choose the next greater
integer: n=4. Substituting n=4 in the equations above, we find two different fc
values: 1.18 MHz and 1.47 MHz. Hence, the cutoff frequency can be selected
as any fc in the range 1.18≤ fc ≤1.47 MHz. Choosing fc =1.3 MHz and the
L1 − C2 − L3 − C4 LPF topology, we find
0.7654 · 50 1.8478
L1 = 6
= 4.68µH and C2 = = 4.52 nF
2π1.3 · 10 2π50 · 1.3 · 106
1.8478 · 50 0.7654
L3 = 6
= 11.3µH and C4 = = 1.87 nF
2π1.3 · 10 2π50 · 1.3 · 106
Example 3
Let us consider the OPAMP second-order low-pass-filter of Fig. 3.32(b) in p. 110
and find the condition for making it a Butterworth filter.
We assume that the source resistance, RS , and the load resistance, RL , are
4RS /RL =1. From Eq. 3.65 in p. 112, we find the transducer power gain of
Eq. 6.2 as
2 2
Vo 1
GT = = (6.18)
Vin 1 + jω(R1 + R2 )C1 − ω 2 R1 R2 C1 C2
or
1
GT = = (6.19)
1+ ω 2 ((R1 + R2 )2 C12 − 2R1 R2 C1 C2 ) + ω 4 (R1 R2 C1 C2 )4
This equation is in the form of Butterworth polynomial of Eq. 6.14, if
(R1 + R2 )2 C1 = 2R1 R2 C2 (6.20)
High-pass-filter
The transfer function of a Butterworth high-pass-filter is written in the following
form
PL 1
= (6.21)
PA 1 + (fc /f )2n
At f = fc , we have PL /PA = 0.5 = −3 dB defining the cut-off frequency of the
high-pass-filter. For f → 0, we have PL /PA → (f /fc )2n defining the asymptote.
We use the following procedure to design an nth-order Butterworth high-
pass-filter of cutoff frequency fc for load and source impedances of R:
6.2. POLYNOMIAL FILTERS 228
1. Use the corresponding Butterworth table value to find the inductor value
as
(1/bi )R
Li = (6.22)
2πfc
2. Use the corresponding Butterworth table value to find the capacitor value
as
1/bi
Ci = (6.23)
2πRfc
Note that we use 1/bi rather than bi for high-pass-filter component values.
Below the cutoff frequency, the filter attenuates the signals with 20n dB/decade
or 6n dB/octave.
As in the case of LPF, the transfer function of an HPF can be drawn eas-
ily using the asymptotic lines. This time, the asymptotic line with a slope
+20n dB/decade passing through (fc , 0 dB) is drawn to indicate the response
below the cutoff frequency.
Example 4
Let us design a fifth-order Butterworth high-pass-filter for a cut-off frequency
of fc =15 MHz for source and load impedances of R=50 Ω. We can use either
L1 − C2 − L3 − C4 − L5 topology or C1 − L2 − C3 − L4 − C5 topology. Again,
the second one (shown in Fig. 6.6(b)) is preferable since it uses smaller number
of inductors. Using the n = 5 values in Table 6.1 and R=50 Ω we find
1/0.618
C1 = C5 = = 343 pF,
2π · 50 · 15 · 106
1/2.0 (1/1.618)50
C3 = 6
= 106 pF, L2 = L4 = = 0.330 µH
2π · 50 · 15 · 10 2π · 15 · 106
This filter attenuates the signal at 7.5 MHz (one octave lower than fc =15 MHz)
by 6×n=6×5=30 dB.
Band-pass-filter
Band-pass-filters have a passband with a center frequency of fo and 3-dB cut-off
frequencies of f1 and f2 . The bandwidth of the filter is ∆f = f2 − f1 . It is
composed of series and parallel LC branches resonating at fo . They have a
transfer function of the following form:
PL 1
= (6.24)
PA 1 + (fo /∆f )2n (f /fo − fo /f )2n
−10
Transfer function (dB)
−20
−30
−40
First−order BPF
−50 Second−order BPF
Third−order BPF
Asymptotes
−60
−1 0 1
10 10 10
f/fo
Example 5
Let us design a third-order band-pass-filter centered at f0 =28 MHz with a band-
width of ∆f =5 MHz for RS = RL =50 Ω. We first design a low-pass-filter with
n=3 using C1 − L2 − C3 topology:
1.0 2.0 · 50
C1 = C3 = = 637 pF and L2 = = 3.20µH
2π · 50 · 5 · 106 2π · 5 · 106
Now we add parallel inductors L1 and L3 and a series capacitor C2 with values
found from the resonance condition at 28 MHz, Eq. 5.4 (page 184);
25330 25330
L1 = L3 = = 0.0510µH = 51 nH, C2 = 2 = 10.0 pF
282 · 637 28 · 3.2
A schematic of this filter is given in Fig. 6.8. We find −3 dB frequencies from
RS C2 L2
+
Vin C1 L1 L3 C3 RL
Eq. 6.25, f1 =25.61 MHz and f2 =30.61 MHz. The filter attenuates input signal
by 20 × 3 = 60 dB at f =50 MHz (one decade higher than ∆f =5 MHz), and at
f =15.68 MHz (one decade lower than fo2 /∆f = 282 /5=156.8 MHz.
n b1 b2 b3 b4 b5 b6 b7 b8
3 0.6292 0.9703 0.6292
4 0.7129 1.200 1.321 0.6476
5 0.7563 1.305 1.577 1.305 0.7563
6 0.7814 1.360 1.690 1.535 1.497 0.7098
7 0.7969 1.392 1.748 1.633 1.748 1.392 0.7969
8 0.8073 1.413 1.782 1.683 1.853 1.619 1.555 0.7334
n b1 b2 b3 b4 b5 b6 b7 b8
3 1.228 1.153 1.228
4 1.303 1.284 1.976 0.8468
5 1.339 1.337 2.166 1.337 1.339
6 1.360 1.363 2.239 1.456 2.097 0.8838
7 1.372 1.378 2.276 1.500 2.276 1.378 1.372
8 1.380 1.388 2.296 1.522 2.341 1.493 2.135 0.8972
T0 (x) = 1
T1 (x) = x
T2 (x) = 2x2 − 1
T3 (x) = 4x3 − 3x
T4 (x) = 8x4 − 8x2 + 1
T5 (x) = 16x5 − 20x3 + 5x
T6 (x) = 32x6 − 48x4 + 18x2 − 1
This filter gives a sharper cutoff (a better rejection outside the passband) than
the Butterworth filter, but it has a ripple in the passband. The design method
is the same as that of the Butterworth filter using the prototype values given
in Tables 6.2 or 6.3. Filter characteristics are shown for several n values in
Figs. 6.9 and 6.10.
Example 6
Let us design a 5.-order 0.2 dB ripple Chebyshev low-pass-filter for a cutoff
frequency of fc =20 MHz for source and load impedances of RS = RL =300 Ω.
We use C1 − L2 − C3 − L4 − C5 topology. Using the n = 3 values in Table 6.3
and R=300 Ω we find
1.339 1.337 · 300
C1 = C5 = 6
= 35.5 pF and L2 = L4 = = 3.19µH
2π300 · 20 · 10 2π20 · 106
6.2. POLYNOMIAL FILTERS 232
-20
Transfer function (dB)
-30
-40
-50
-60
1 1.5 2 2.5 3 3.5 4 4.5 5
f/f c
Figure 6.9: Chebyshev LPF response with 0.01 dB ripple for different n values.
At f = fc , the filters have 0.01 dB loss.
2.166
C3 =
= 57.5 pF
2π300 · 20 · 106
At f = fc , the filter has an attenuation of 0.2 dB. We can estimate the perfor-
mance of the filter at 40 MHz using Fig. 6.10: Since fc =20 MHz, at 40 MHz
(f /fc = 2) the signal is attenuated by 38 dB (voltage is 0.013 times).
-20
Transfer function (dB)
-30
-40
-50
-60
1 1.5 2 2.5 3 3.5 4 4.5 5
f/f c
Figure 6.10: Chebyshev LPF response with 0.2 dB ripple for different n values.
At f = fc , the filters have 0.2 dB loss.
n1 : n2 n1 : n2
C Lp RL Lp RL
Example 7
Design a transformer to transform 20 Ω into 50 Ω at fo =10 MHz. We have a
core with AL =3 nH/turns2 .
Referring to Eq. 6.30, we should choose the turns ratio as
r r
n1 R 50
= = = 1.58
n2 RL 20
Let us choose n1 : n2 =11:7, since 11/7=1.57. The inductance of the primary
is Lp = n21 AL =363 nH. We choose a capacitance to tune out the inductance at
10 MHz. Using the resonance formula of Eq. 5.4 on page 184:
25330
C= = 698 pF
102 · 0.363
6.3. IMPEDANCE MATCHING 235
Example 8
Design a transformer to transform the 4 Ω speaker impedance of the heavy-metal
concert given on page 210 to 8 Ω to recover the lost 111 W. We need a wideband
transformation covering the whole audio range (20 Hz to 20 kHz). Hence the
lowest frequency of interest is f1 =20 Hz. We have a core with AL =650 nH/T2 .
Referring to Fig. 6.11(b), we should choose the turns ratio of the transformer
from Eq. 6.33 as r r
n1 R 8
= = = 1.41
n2 RL 4
To have a wide-band transformer, from Eq. 6.32 we should choose
2
n1 RL 8 4
Lp ≥ 4 =4 = 256 mH
n2 ω1 4 2π20
p
Hence n1 = 256000/0.65=628. We find n2 =445.
6.3. IMPEDANCE MATCHING 236
Example 9
Design a transformer to transform 10 Ω into 50 Ω between 4 MHz to 15 MHz.
We have a core with AL =3 nH/turns2 .
Referring to Fig. 6.11(b), we should choose the turns ratio as
r
n1 50
= = 2.24
n2 10
To have a wide-band transformer, from Eq. 6.32 we should choose
50 10
Lp ≥ 4 = 7.96 µH
10 2π4 · 106
p
Let Lp =8 µH>7.96 µH, hence n1 = 8000/3=51.6. Choose n1 =52 turns. We
find n2 =23.2. Since n2 is not close to an integer, let us choose the better pair
of 56:25 (56/25=2.24).
LS LS
RP C RS RP LP RS C RP
C
Example 10
Design a resonant matching circuit to convert 10 Ω to 50 Ω at fo =10 MHz.
Referring to Fig. 6.12(a), RS =10 and RP =50. Hence
RP 50
Q2 + 1 = = =5
RS 10
jX jX jX
(a) (b)
1 X2
ZI = jX + = (6.37)
1/(−jX) + 1/(jX + Z) Z
6.3. IMPEDANCE MATCHING 238
This functional circuit is used for many matching and filtering purposes. For
example, it can convert a series resonant circuit into a parallel resonant circuit.
Assume that Z is the impedance of a series resonant RLC circuit:
1
Z = jωL + R + (6.38)
jωC
ZI becomes
X2
ZI = 1 (6.39)
jωL + R + jωC
Example 11
Design an impedance inverter to convert Z=5 Ω to ZI =50 Ω at 28 MHz.
From Eq. 6.37, we have X 2 =ZI Z=50·5=250 or X=15.8 Ω. We can use
either of the inverters shown in Fig. 6.13. To generate +jX we use an inductor
of value 15.8/(2π28 · 106 )=89.8 nH. To generate −jX, we use a capacitor of
value 1/(15.8 · 2π28 · 106 )=359 pF.
Example 12
Let us transform a resistance of RL =50 Ω to R=100 Ω at a frequency of 16 MHz
using different techniques. p √
Using a transformer we need a turns ratio of n1 /n2 = 100/50 = 2. If
we can make the primary magnetizing inductance large enough compared to
R = 100Ω, it is a wide-band transformation (as in Fig. 6.11(b)). Setting ωLp =
4R = 400, we find Lp = 3.97µH. Using a toroidal core with AL =4.5 nH/turns2 ,
we need n1 =30 turns, which is a reasonable number. Hence n2 =21 turns.
Alternatively, we can make a resonant transformer with less number of
turns: Choose n1 =14 turns and n2 =10 turns. In this case, Lp = AL n21 =
4.5 · 142 =882 nH. To tune out the primary inductance, we need a capacitor of
value C = 25330/(162 · 0.882)=112 pF in parallel with it (as in Fig. 6.11(a)).
As a third method, we can use an L-section as in Fig. 6.12. We choose
RS = 50Ω, and we need to have RP = 100Ω. So, we have Q2 + 1 = RP /RS = 2,
and we find Q = 1. Since Q = ωLS /RS = 1, we determine LS =497 nH
and LP =994 nH. The capacitor of the L-section is found from the resonance
condition: C = 25330/(162 · 0.994)=99 pF.
As a fourth method,
√ we√ use one of the impedance inverters of Fig. 6.13.
We choose X = RL R = 50 · 100=70.7. Hence, the series inductor to gen-
erate +jX is L = 70.7/ω=703 nH and the shunt capacitor to generate −jX is
C=140 pF.
6.3. IMPEDANCE MATCHING 239
ZI Zp
C2 L2
-jX
L1 C1 R ZI L1 C1 R
jX jX
X=R
(a) (b)
Ce
R / ωo
C1 1/Rωo
L1 L1 C1 R Le C1 Le C1 R
R / ωo
(c) (d)
1.4142 1.4142 · R 1 1
C1 = , L2 = , L1 = 2 , C2 = 2
∆ωR ∆ω ωo C1 ωo L 2
Now, consider the impedance seen at the input of inverter of Fig. 6.13(b) with
X = R shown in Fig. 6.14(b). We can write
X2 R2
2 1 1 1
ZI = = =R + jωC1 + = + jωL2 + R
Zp Zp jωL1 R jωC2
−10
Transfer function (dB)
−20
−30
−40
−60
−1 0 1
10 10 10
f/fo
Example 13
Let us design a second-order band-pass-filter (with no series branches) centered
at fo =16 MHz with a bandwidth of ∆f =1 MHz for RS = RL =50 Ω. We
first design a low-pass-filter with n=2 and bandwidth of 1 MHz using C1 − L2
topology:
1.41
C1 = = 4.5 nF
2π · 50 · 1 · 106
Now we add a parallel inductor L1 to C1 from the resonance condition at
16 MHz, using the resonance formula of Eq. 5.4 (page 184):
25330
L1 = = 0.22µH = 220 nH
162 · 4500
To get rid of the series branch, we use an inverter with X=50 Ω. We need
an inductor of L = R/ωo = 50/(2π16 · 106 ) =497 nH, and a capacitance of
Ce =199 pF. Since Le is the parallel combination of two inductors
RL1 220 · 497
Le = = = 152 nH (6.42)
R + ωo L1 220 + 497
Hence we determined all the values in Fig. 6.14(d).
6.4. CRYSTAL FILTERS 241
Quartz crystals
Quartz crystals have a very high mechanical Q, in the order of 100,000, due to
their orderly single-crystal structure. A poly-crystalline material does not have
such a high mechanical Q. Since quartz is also piezoelectric, the mechanical
resonance directly influences the electrical properties, and hence quartz displays
a similarly high electrical Q.
The schematic symbol of a crystal is given in Fig. 6.16(a). There are many
rs
Co Ls
Cs
(a) (b)
Figure 6.16: Quartz crystal (a) symbol and (b) equivalent circuit.
frequency and its overtones. We are interested in the fundamental resonance fre-
quency of only one mode of vibration. The equivalent circuit of a quartz crystal
in the vicinity of this frequency is also given in Fig. 6.16(b). These equivalent
circuit models the impedance at the (electrical) terminals of the quartz crystal.
The mechanical properties of quartz and the dimensions of the plate, determine
all three of the series circuit elements. Ls is proportional to the mass of the
plate, and Cs is determined by the compliance (inverse of stiffness) of quartz
crystal. Resistance rs models the friction losses during vibration. The only
inherently electrical component is Co , which is the capacitance between the
electrodes of quartz crystal.
The model in Fig. 6.16(b) has two resonance frequencies:
1
fs = √ (6.43)
2π Ls Cs
and
1
fp = p . (6.44)
2π Ls Cs Co /(Cs + Co )
fs is the resonance frequency of the series rs Ls Cs branch and the impedance
decreases down to rs at this frequency (in parallel with Co ). fp , on the other
hand, is the parallel resonance frequency, where the inductance Ls resonates
with the series combination of Cs and Co , i.e., Cs Co /(Cs + Co ). We note that
fs < fp .
Ls ranges from a few mH to over 10 mH for quartz crystals at 15 MHz. Cs
is in the order of fF (femto Farad, 10−15 F) and rs is in the range of a few ohms
to a few tens of ohms. Co is an electrical component, the clamp capacitance,
usually a few pF. We are interested in the frequencies in the vicinity of series
resonance.
R1 jX jX Vo
impedance inverter
(a)
R1 Vo Ro =R1+rs L1 Vo
Ls Cs rs 2 Ls Ro =
Vin CsX Ls/X2 X2/(R2+rs) Vin
XT1 C2 Ls/X2 X2/(R2+rs)
(b) (c)
Figure 6.17: (a) Two-crystal ladder filter, (b) the equivalent circuit after inver-
sion operation, (c) the low-pass-filter prototype.
R1 Li Li Vo R1 Vo
Vin X1 Ci X2 R2 Vin X1 Ci X2 R2
impedance inverter
(a) (b)
Figure 6.18: Impedance inverter for two-crystal BPF, (a) Proper inverter, (b)
Actual circuit with the inductors ignored.
Example 14
Suppose we would like to design a second-order BPF centered at fo =16 MHz
with a bandwidth of ∆f =10 kHz. If we design this filter using inductors and
capacitors, we end up with a very large inductor and a tiny capacitor in the series
branch, and a tiny inductor and a very large capacitor in the shunt branch. Let
us use two series 16 MHz crystals with an inverter in between. The crystal has
Ls =15 mH and rs = 15Ω. From Eq. 6.48, we find
2π · 104 · 15 · 10−3
X = Ro = R1 + rs = R2 + rs = = 666Ω or R1 = R2 = 651Ω
1.4142
The inverter inductance of value Li =6.6 µH is ignored in comparison to
Ls =15 mH. The inverter capacitor is given by
1 1
Ci = = = 14.9 pF
2πfo X 2π · 16 · 106 · 666
We can reduce the bandwidth ∆f of the filter by a factor of two if we choose
R1 = R2 = 318Ω and Ci =29.8 pF.
Example 15
Design a third-order BPF centered at fo =8.00MHz with a bandwidth of
∆f =20 kHz. We have quartz crystals with fs =8 MHz, rs =10 Ω and
Q=100,000.
For the third-order BPF, we need three crystals and two inverters, as shown
in Fig. 6.19. From quartz crystal data, we find Ls = Qrs /ωs = 19.9 mH. For
R1 Li Li Li Li Vo
Vin X1 Ci X2 Ci X3 R2
n=3, the prototype filter values are 1, 2, 1. To match the inductance of the
crystal X1 to the first inductance, L1 , of the LPF we should have
Ro Ro
Ls = 19.9 mH = L1 = = (6.51)
2π∆f 2π20 · 103
Hence Ro =2.5 kΩ. To match the inductance of the crystal X2 to the capacitance,
C2 , of the bandpass filter, we have
2 2 Ls
C2 = = 3
= 6.37 nF = 2 (6.52)
2π∆f Ro 2π20 · 10 · 2500 X
√
Hence X = Ro / 2=1770 Ω. Hence Li = 35 µH, and Ci =11.2 pF. Note that
the third inductance of the BPF, L3 = L1 , is matched to Ls automatically by
going through two inverters. We can ignore Li , since it is small compared to Ls .
The termination resistors should be selected as R1 = R2 = Ro − rs = 2490 Ω.
6.6 Examples
Example 16
We have a source with RS = 50 Ω and a load impedance as given in Fig. 6.20(a).
Design a matching circuit to maximize the power transfer to the 2 Ω load resistor
for ω = 2π28 · 106 . We have a core with AL = 2.3 nH/T2 .
L
+ 50Ω Rs + 50Ω Rs
? 2Ω 2Ω
sin(ωt) 500pF sin(ωt) C1 C2 500pF
(a) (b)
L2 n1 : n2
+ 50Ω Rs + 50Ω Rs
Lp 2Ω 2Ω
sin(ωt) C3 500pF sin(ωt) C3 500pF
(c) (d)
Figure 6.20: (a) Source and load impedances for Example 16, (b) an impedance
inverter used as a matching circuit, (c) an L-section used as a matching circuit,
(d) a transformer used as a matching circuit.
Solution
Many possibilities for the matching circuit exist. As a first method, let us
use an impedance inverter. We prefer to use a “π” type impedance inverter
as depicted in Fig. 6.20(b), since there is already a shunt capacitor as part of
the load impedance. We find X 2 = 50 · 2 = 100, or X = 10. Hence C1 =
1/(10 · 2π28 · 106 ) = 568 pF and L = 10/(2π28 · 106 ) = 56.8 nH obtained with
5 turns on the core. We also have C2 + 500 = 568, hence C2 = 68 pF.
As a second method, let us utilize an L-section. First, we tune out the
capacitor with a parallel inductor of value
25330
Lp = = 65 nH
282 · 500
This can be obtained with (squeezed) 5 turns. We need an L-section with
Q2 + 1 = 50/2 = 25 or Q = 4.9. Since Q = ωL2 /Rs , we get L2 = 55.7 nH (5
turns) and C3 = 560 pF.
As a third method, we use a transformer whose secondary inductance tunes
out the 500 pF capacitor. The turns ratio of the transformer must be
r
n1 50
= =5
n2 2
Hence the primary has 25 turns, and the secondary has 5 turns.
6.6. EXAMPLES 247
Example 17
Design a band-pass-filter between the source and load resistor of 500 Ω. 3-dB
frequencies are 1 kHz to 5 kHz. We would like to reject the 500 Hz and 9.5 kHz
by at least 20 dB.
Solution
Since ∆f = 5000 − 1000 = 4000, using Eq. 6.25 we write
r r
∆f 2 ∆f 40002 4000
f1 = 1000 = fo2 + − = fo2 + −
4 2 4 2
and find fo = 2236 Hz. Since −20 dB means a power ratio of 0.01, using Eq. 6.24
for f = 500 Hz, we get
PL 1
0.01 > =
PA 1 + (2236/4000)2n (500/2236 − 2236/500)2n
To find the order of the filter, we try different n values: For n = 2, PL /PA = 0.03
and for n = 3, PL /PA = 0.006 < 0.01. For f = 9.5 kHz, we get
PL 1
0.01 > =
PA 1 + (2236/4000) (9500/2236 − 2236/9500)2n
2n
Example 18
The OPAMP circuit shown in Fig. 6.21 is a first-order band-pass-filter. Find the
center frequency and the bandwidth in terms of the given component values.
R1 C
V3 V2 − R
Vin
R2 V1 Vo
+
Solution
Assuming that the OPAMP is not saturated, we have V1 = V2 = 0. We can
write the node equation for V3 as
V3 − Vin V3 V3 − Vo V3
+ + + =0
R1 R2 1/jωC 1/jωC
Writing KCL at V2
−V3 −Vo Vo
+ = 0 or V3 = −
1/jωC R jωRC
R1 + R2 − ω 2 R1 R2 RC 2 = ±ω 2 R1 R2 C
Example 19
The OPAMP circuit shown in Fig. 6.22 is a band-stop-filter or a notch filter.
It can be used to eliminate unwanted frequencies in a signal. Find the notch
frequency and the 3-dB bandwidth of the filter.
6.6. EXAMPLES 249
R V3 R
2C V2 −
Vin V1 Vo
+
C V4 C
R/2
Solution
We write the node equations for V3 , V4 and V1 as
V3 − Vin V3 V3 − V1
+ + =0 (6.53)
R 1/jω2C R
V4 − Vin V4 V4 − V1
+ + =0 (6.54)
1/jωC R/2 1/jωC
V1 − V3 V1 − V4 1+X 1
+ =0 or V4 = V1 − V3 (6.55)
R 1/jωC X X
where we used X = jωRC to simplify the notation. From Eq. 6.53, we get
X
V3 = (Vin + V1 ) (6.56)
2 + 2X
Combining Eqs. 6.54 and 6.55, we find
X2 2 + 4X + X 2
V3 = − Vin + V1 (6.57)
2 + 2X 2 + 2X
Equating Eqs. 6.56 and 6.57, we get
Vo V1 1 + X2 1 − ω 2 R2 C 2
= = 2
= (6.59)
Vin Vin 1 + 4X + X (1 − ω 2 R2 C 2 ) + jω4RC
since the OPAMP is configured as a unity gain buffer and Vo = V1 . The transfer
function approaches unity when ω → 0 and when ω → ∞. The notch occurs
when the numerator is zero:
1
ωo = (6.60)
RC
To√find the 3-dB frequencies, we set the magnitude of the transfer function to
1/ 2:
2
Vo (1 − ω 2 R2 C 2 )2 1
= = (6.61)
Vin (1 − ω 2 R2 C 2 )2 + (ω4RC)2 2
6.6. EXAMPLES 250
Example 20
Consider the circuit of Fig. 6.23(a) where the reactances are specified at f =
20 MHz. Find the input impedance Zin at the same frequency, f , using the
impedance inverter formula, rather than series/parallel combination of impedances.
(a) (b)
Figure 6.23: (a) Circuit for Example 20, (b) Modified circuit for simple analysis.
Solution
To make the circuit look like an impedance inverter we separate the rightmost
inductor into two inductors or two reactances j50 = j40 + j10, as shown in
Fig. 6.23(b). Hence the effective load impedance of the inverter becomes Zout =
12 + j10. We can find the input impedance, Zin , using the impedance inverter
formula with X = 40:
X2 402
Zin = = = 78.7 − 65.5
Zout 12 + j10
Example 21
Find Zin for the circuit of Fig. 6.23(a) at 22 MHz.
Solution
The circuit at 22 MHz becomes as shown in Fig. 6.24(a) since the inductive
reactances increase by 10% and capacitive reactances decrease by 10%. We
modify the circuit so that a portion of the circuit becomes an inverter as in
Fig. 6.24(b). We can find the input impedance at 22 MHz using an impedance
inverter with X=36.4 and a series inductive reactance.
36.42
Zin = j7.6 + = 32.4 − j42.6
12 + j18.6
6.6. EXAMPLES 251
(a) (b)
Figure 6.24: (a) Circuit for Example 21, (b) Modified circuit for simple analysis.
6.7. PROBLEMS 252
6.7 Problems
1. Find the amplitude and phase of the voltage at the output of the circuit
given in Fig. 6.25 when the input voltage is 2 cos(2π16 · 106 t) volts. Find
the amplitude of the current flowing through the capacitor.
3. Find the amplitude and phase of the voltage at the output of the circuit
given in Fig. 6.26 when the input current is 10 cos(1.82 · 108 t) mA. Find
the amplitude of the current flowing through the inductor.
4. Find the 3-dB cutoff frequency of the low-pass-filter given in Fig. 6.27.
16. Find the input impedance, Zin , of the circuit in Fig. 6.29 using the
impedance inverter formula.
Zin 10
-j60 -j60 -j30
DIODES IN
TELECOMMUNICATIONS
D D
vhw(t) vo(t)
+ +
vIF(t) R vIF(t) R C
(a) (b)
The waveform of this signal is plotted in Fig. 7.2(a) for Vc = 1, Vm = 0.7, and
fIF =20 kHz, fm = 1 kHz. The modulation index (see Eq. 1.6 at page 7) of this
signal is m = Vm /Vc =0.7.
AM demodulation is the act of separating the information signal vm (t) from
its carrier Vc cos(2πfIF t). The simplest and oldest method of doing this is called
envelope detection. In an envelope detector, the signal is half-wave rectified and
then low pass filtered. We employ the rectification property of diodes in envelope
detection. A half-wave rectifier is shown in Fig. 7.1(a). It is composed of an
AM signal
2
vIF(t)
0
−1
−2
0 0.5 1 1.5 2
Time x 10
−3
1
vhw(t)
−1
−2
0 0.5 1 1.5 2
Time x 10
−3
1
vo(t)
−1
−2
0 0.5 1 1.5 2
Time −3
x 10
Figure 7.2: AM Waveform, vIF (t) (upper) for fIF =20 kHz, fm = 1 kHz, half-
wave rectified AM waveform, vhw (t) (middle) and envelope detector output,
vo (t) (lower).
Otherwise, the detected envelope signal suffers from what is known as failure
to follow distortion or diagonal distortion. This upper limit on RC causes some
ripple on the detected waveform, particularly during the up-sloping phases of
the envelope, as seen in Fig. 7.2(c).
We used a very low carrier frequency (fIF = 20 kHz) to demonstrate the
function of the envelope detector and to exaggerate the ripple. In a real case,
the carrier frequency is much higher. Fig. 7.3 shows vhw (t) and vo (t) for fIF =
100 kHz. In this case, the ripple is smaller, and the envelope detector output
is approximately the same as the original sine wave of vm (t) except for the DC
shift. The ripple is negligible with a carrier frequency in the MHz range.
The highest ripple occurs when the AM signal amplitude is maximum (Vc +
Vm ). If fm ≪ fIF and 1/RC ≪ fIF , the peak-to-peak ripple in the output can
be estimated from the peak-to-peak ripple formula of Eq. 4.7 for the half-wave
rectifier:
1
Vr = (Vc + Vm ) (7.3)
RCfIF
where the voltage drop across the envelope detector diode is ignored.
Note that the DC shift amount (or the average value) at the envelope detec-
tor output is the same as the amplitude of the carrier, Vc . The DC shift voltage
can be used to determine the carrier amplitude of the AM signal.
Half−wave rectified AM signal
2
1
vhw(t)
−1
−2
0 0.5 1 1.5 2
Time x 10
−3
1
vo(t)
−1
−2
0 0.5 1 1.5 2
Time −3
x 10
Figure 7.3: Half-wave rectified AM waveform, vhw (t), (upper) for fIF =100 kHz,
fm = 1 kHz and envelope detector output, vo (t) (lower).
the detected envelope of a vIF (t) of, for example, 2 Vpp amplitude is severely
distorted, as illustrated in Fig. 7.4(b).
Half−wave rectified AM signal with Vo=0.6
2
vhw(t) 0
−1
−2
0 0.5 1 1.5 2
Time x 10
−3
1
vo(t)
−1
−2
0 0.5 1 1.5 2
Time −3
x 10
Figure 7.4: Half-wave rectified AM waveform, vhw (t), (upper) for fIF =100 kHz,
fm = 1 kHz for a real diode with Vo =0.6 V and the corresponding envelope
detector output, vo (t) (lower).
CB CB
Vdc
Idc RB
Idc Idc
L
Cc Cc vo(t)
vo(t) +
+ +
D Vo D
vIF(t) R C vIF(t) R C
(a) (b)
is a very slow acting diode. Once the diode is turned on, it does not turn off
quickly. Conversely, it does not turn on easily if the diode is off. Although it is
a very slow diode, it has a very small junction capacitance making it suitable for
high frequencies. (p-n junction rectifier diodes like 1N4007 are also very slow,
but they have a very large junction capacitance.)
PIN diode acts like a regular diode for low-frequency signals (for signals lower
than about 100 kHz) with highly nonlinear characteristics. At high frequencies
(higher than about 5 MHz), it acts like a linear resistor even for large signal
amplitudes. This high-frequency resistance is inversely proportional to the DC
or low-frequency current passing through the diode. If the DC current is not
present, the PIN diode acts as an open-circuit at high frequencies even if the
instantaneous high-frequency voltage across the diode is a large positive voltage.
A short burst of high-frequency positive voltage is not enough to turn on the
diode.
On the other hand, the PIN diode acts as a resistor with a small value (a
few ohms), when there is a positive DC current through it. This is true even if
the instantaneous high-frequency voltage may be several volts negative.
With these interesting properties, a PIN diode can be used as the element
we need for the automatic gain control mechanism.
Vdc
R
R45 C53 D50
+ IF amp vo(t)
vIF(t) L51 C55
C81
♦ TRC-11 has one signal diode in the amplitude demodulator, one PIN diode
to change the gain of the IF amplifier, and LED as the signal presence
indicator, and one LED as power indicator.
7.4. PROBLEMS 262
7.4 Problems
1. The AM signal 2[1 + 0.7 cos(2π103 t)] cos(2π28 · 106 ) is fed to an envelope
detector as in Fig. 7.1(b) with R=1 kΩ. Find the maximum value of
capacitor, to avoid failure to follow distortion.
4. In Fig. 7.8, v(t) is 0.5 cos(ωt) for both circuits. Assume that the diode can
be modelled by the approximate model of Fig. 4.3(c) and (d) on page 136,
with V0 =0.7 V. What is vout (t) for both circuits, if I = 0? Find the
minimum value of I for which there is an undistorted replica of v(t) at the
output. For both circuits, find vout (t) for this value of I. What is the value
of I such that time varying part of vout (t) is exactly half wave rectified (but
scaled, of course) form of v(t)? (Hint: First find the Thévenin equivalent
circuit, comprising both sources, across the detector circuit)
5. Consider the 28 MHz amplifier shown in Fig. 7.9 built using an OPAMP.
RF resistance of the PIN diode can be varied between 10 Ω to 1000 Ω,
while the DC current source IDC is varied between 10 mA to 0.1 mA.
7.4. PROBLEMS 263
Find the gain, |Vout /Vin |, of the amplifier at 28 MHz, for IDC =10 mA
and 0.1 mA. Assume that the reactance of 1 nF capacitor is negligible at
28 MHz.
Vin +
Vout
−
1nF 51
510
IDC 470
PIN
6. Find the value of the resistor in the LED drive circuit of Fig. 4.17(c) for a
voltage source of Vs =5 V, when a white-light power LED with a current
of 1.5 A is to be connected. With 1.5 A current, the voltage drop across
the LED is 3.1 V. If the light conversion efficiency of the LED itself is
100 lumens/W, what is the light conversion efficiency of the LED with
the series resistor?
Chapter 8
FREQUENCY
CONVERSION
8.1 Mixers
As discussed in Chapter 1, mixers are nonlinear devices with three ports to
perform the multiplication process. They can be built from diodes, transistors
or other nonlinear devices. The concept of mixing is easily understood with a
switch mixer.
The switch is opened and closed at the frequency of vLO . The switch is
closed if vLO (t) ≥ 0, and it is open if vLO (t) < 0. Therefore, we can write the
fLO vLO(t) LO
Switch
RF vIF(t)
vRF(t) IF
fRF
output signal, vIF (t), in terms of the input signal, vRF (t), and the switching
function, s(t), as
vIF (t) = vRF (t)s(t) (8.2)
Supply Bypass
Cap
+ Modulator
vm(t)=Vm cos ωmt
Amplitude
Tuned Modulated
RF
+ Amp Output
vc1(t)=Vc1cos ωct
Figure 8.2: Amplitude modulator circuit. The input signal of the tuned RF
amplifier should be so large that the output magnitude is determined by the
supply voltage.
the amplified signal follows the variation (see Fig. 8.2. This principle is used in
TRC-11 to obtain an amplitude modulator.
The supply voltage modulation is achieved using a PNP BJT circuit, the
input of which is controlled by the modulation signal at audio frequency. The
collector of the BJT is connected to the supply voltage of an RF amplifier
through an RF bypass capacitor. The bypass capacitor selected such that it
acts like a short circuit at RF frequency but it is an open circuit at the audio
frequencies. Hence the supply voltage of the tuned RF amplifier varies exactly
like the modulation signal.
8.3 Oscillators
An oscillator is a sinusoidal signal generator. Frequency is determined by a
resonant circuit inside the oscillator. The resonator can be built, for example,
from LC circuits or quartz crystals.
Every resonator has some loss. If this loss can be compensated by an active
circuit, a continuous sinusoidal signal can be obtained. Depending on the oscil-
lator structure, some oscillators provide a square-wave output, rather than the
sinusoidal signal.
It is possible to obtain an oscillator by using a positive feedback between the
output and input of an amplifier using a resonator. The resonator determines
the frequency of oscillation while the amplifier provides the continuity of the
oscillation.
The quality of an oscillator is determined by the purity and stability of the
output signal. Oscillators using high-Q resonators usually have high quality out-
put signals. Oscillators using LC circuits as their resonators typically produce
low quality signals, since Q of electrical circuits are limited. Oscillators using
quartz crystals are called crystal oscillators and they provide a well-defined and
stable frequency output signal. Many laboratory signal generators use crystal
oscillators as their reference frequency. If higher stability is desired, crystals
may be placed in temperature-controlled ovens to get frequencies that are more
precise.
8.3. OSCILLATORS 267
v1(t)
v1=V1cos(ωot)
(a) i(t) L C R1
V+ v2(t)
v1(t) + V+
v2(t)=Av1(t)
A
(b) i(t) L C R1 −
V- V-
R2
V+ v2(t)
v1(t) + V+
v2(t)
A
(c) i(t) L C R1 −
V- V-
R2
V+ v2(t)
v1(t) + V+
v2(t)
A
(d) L C R1 −
V-
V-
Figure 8.3: Oscillator concept: (a) RLC circuit driven by a wide-band current
source, (b) amplification, (c) positive feedback, (d) current source removed.
sinusoidal components at all frequencies and its amplitude is very small. The
noise in electronics
√ is such a signal. Only the current component in the vicinity
of ωo = 1/ LC generates a voltage v1 (t) across the tank circuit. The amplitude
of v1 (t) is very small also. We expect to observe v1 (t) as
At this stage, let us assume that the gain is not large enough to saturate
the amplifier. When a feedback path to the positive input is provided by means
of a resistor R2 as shown in Fig. 8.3(c), v1 (t) is modified. Initially, an additive
sample from output increases v1 (t) to
R1
v1 (t) = V1 cos(ωo t) + AV1 cos(ωo t) (8.9)
R1 + R2
The additive component has an amplitude of R1 /(R1 + R2 )(AV1 ) which is
much larger than the signal directly created by i(t). This component alone
drives the amplifier deep into saturation, since AR1 /(R1 + R2 )(AV1 ) ≫ V + .
We immediately have a square wave at the output as v2 (t). The peak-to-peak
amplitude of this wave is 2V + . This waveform is also shown in Fig. 8.3(c).
When we have a square wave at the output, the feedback signal is also a
square wave. However, the tank circuit picks the fundamental component of
this square wave, producing a v1 (t) as
R1
v1 (t) = V1 cos(ωo t) + b1 V + cos(ωo t) (8.10)
R1 + R2
From Eq. 1.4 of page 5 we know that b1 = 4/π is the coefficient of the funda-
mental component in a square wave. Since the output of the amplifier cannot
change any more, the circuit operation is stabilized with a square wave at its
output. The tank circuit determines the frequency of this signal.
The input signal v1 (t) is predominantly the feedback signal. If we remove
the current source from the circuit, the output is still a square wave. Neither
v2 (t) nor v1 (t) are affected, as shown in Fig. 8.3(d).
Practically, we never include an explicit current source unit to start the os-
cillation, because it is not necessary. There is always noise in electronic circuits,
creating the current needed for the start of the oscillation. The current source
is always there.
It is possible to deduce from the above discussion that if
R1
A=1 (8.11)
R1 + R2
we have a sustained oscillation, once it starts. In this case, the output waveform
is sinusoidal and amplifier works in linear region all the time. This condition,
i.e., the product of amplifier gain and the feedback ratio being unity is called
the Barkhausen oscillation criterion.
In the circuit of Fig. 8.3, we have the amplitude limiting mechanism of
saturating amplifier. Since the fundamental component of the saturated output
is b1 V + , the peak-to-peak amplitude, V1pp , of sinusoidal signal v1 (t) is always
R1 8 R1
V1pp = 2 b1 V + = V+ (8.12)
R1 + R2 π R1 + R2
As long as the gain is large enough to keep the amplifier in saturation with
this input, the oscillation is sustained. The larger gain of the amplifier helps
oscillations to start easily. To make sure that the oscillation starts, the feedback
ratio R1 /(R1 + R2 ) and the gain A must be such that
R1
A>1 (8.13)
R1 + R2
8.3. OSCILLATORS 269
Example 1
Suppose we would like to design an oscillator as in Fig. 8.3(d) at fo =12 MHz
using LM7171, an OPAMP with a gain-bandwidth product of GBW =200 MHz.
We have supply voltages of V + =15 V and V − = −15 V.
The gain A of the OPAMP at 12 MHz is GBW/fo =16.6. Using Eq. 8.13,
we write
R1 1
> = 0.06
R1 + R2 16.6
LM7171 data sheet recommends a feedback resistor R2 =510 Ω. Therefore, we
need to choose R1 > 32.5 Ω. Let us choose R1 =330 Ω (comfortably larger than
the limit). To get a Q=10, we choose
R1 ∥ R2 R1 ∥ R2 200
Q= = 10 or L = = = 265 nH
ωo L 2πfo 10 2π · 12 · 106 · 10
From the resonance condition at 12 MHz, we find the capacitor value as
25330
C= = 664 pF
122 · 0.265
From Eq. 8.12, the peak-to-peak signal amplitude, V1pp is given by
8 R1 8 330
V1pp = V+ = 15 = 15
π R1 + R2 π 330 + 510
In the real circuit, V1pp will be less than 15 Vpp , since the saturation voltages
of the LM7171 are about ±13 V rather than ±15 V.
R3 R3
Zin
Iin V1 + V1 +
+
Vin V2 Vo V2 Vo
− L R −
Zin C
R1 R1
R2 R2
(a) (b)
Figure 8.4: (a) OPAMP circuit with a negative input resistance, (b) an oscillator
built with the negative resistance.
−R2
R ∥ Zin = R ∥ (−R) = ⇒∞ (8.18)
R−R
In this case, an oscillation at the resonance frequency of LC can be maintained
indefinitely. In practice, it is difficult to satisfy the precise equality of a negative
resistance to a positive resistance. To guarantee an oscillation, we choose the
values such that
R1 R3
<R (8.19)
R2
There are many subtle subjects in the theory of oscillators, such as frequency
and amplitude stability, phase jitter, etc. We leave these topics to advanced
texts.
Ib Ve Vb I Ve
Vb +
b
+
Z1 Vbe gmVbe L C1 Vbe gmVbe C2
Z2
- -
Zin Ib Ib
(a) (b)
Vb = (gm Z1 Z2 + Z1 + Z2 ) Ib (8.24)
Example 2
Consider a Colpitts oscillator (as in Fig. 8.5(b)) to operate at 28 MHz. If
C1 =33 pF, C2 =33 pF, and gm =0.005 S, find the value and minimum Q of the
inductor.
From Eq. 8.26, we find Zin = −148 − j344. Hence we must have L =
344/(2π28 · 106 ) = 1.96 µH. If the series resistance of L satisfies r ≤ 148 Ω, the
oscillation will be maintained. Hence we should have Q ≥ ωL/r = 2.32.
Rf
Rf Ro
vi vo =
- Avi
X rs Ls Cs
C1 C2 C1 C2
Co
(a) (b)
Figure 8.6: (a) Pierce oscillator using a digital inverter with crystal feedback,
(b) its equivalent circuit.
Example 3
Consider the Pierce oscillator built using a digital CMOS inverter integrated
circuit, 74HC04, and a 16 MHz quartz crystal. The 14-pin integrated cir-
cuit has six inverters inside. The oscillator uses only one of the inverters,
which has a propagation delay of 6 ns with a supply voltage of 6 V. We have
C1 =C2 =12 pF, Rf =1 MΩ and Ro =100 Ω. The 16 MHz quartz crystal has
parameters Ls =15 mH, rs = 15 Ω, Cs =6.6 fF, Co =5 pF.
The amplitude and phase of the transfer function, vi /vo , of the feedback
circuit is plotted in Fig. 8.7. The series and parallel resonant frequencies are
about 15.999 MHz and 16.006 MHz, respectively. 6 ns delay of the inverter
corresponds to a phase shift of (6 ns)(16 MHz)(360o )=34.6o . Hence a phase
shift of −180o − 34.6o = −214.6o is already provided by the inverter. We need
to find the frequency where the feedback network has −360o − (−214.6o ) =
−145.4o = −2.54 rad phase shift. We see that this feedback circuit provides
a phase shift of −2.54 rad between fs and fp near 15.9992 MHz, where the
amplitude is 12.2 dB. Since the gain of the inverting amplifier is definitely larger
than −12.2 dB, the oscillation condition is easily satisfied at this frequency. We
note that it is possible to shift the oscillation frequency slightly by changing the
values of C1 and C2 .
20 0
10 - /6
0 -2 /6
Phase of v i/v o
|v i/v o | (dB)
-10 -3 /6
-20 -4 /6
-30 -5 /6
-40 -
-50 -7 /6
15.995
15.996
15.997
15.998
15.999
16.001
16.002
16.003
16.004
16.005
16.006
16.007
16.008
16.009
16.01
16
f (MHz)
Figure 8.7: The amplitude and phase of the transfer function, vi /vo , of the
feedback circuit of Example 3 Pierce oscillator.
♦ There are two oscillators in TRC-11. The first one is the local oscillator
that generates a signal at 12.00 MHz, which is a Colpitts crystal oscil-
lator as shown in Fig. 8.8. Here, SA602A is an integrated circuit acting
as a transconductance amplifier necessary for oscillation. We have the
capacitors of Colpitts oscillator as C1 = 39 pF and C2 = 18 pF.
8.3. OSCILLATORS 274
VCC
100pF Osc
OSCA
Out
SA602A
10pF 18pF
OSCB
12MHz
GND
Figure 8.9: (a) The symbol of a varactor diode, (b) the capacitance variation
of a varactor diode as a function of reverse voltage, (c) a voltage controlled
oscillator built by a varactor.
ode depends on the level of reverse bias voltage. The variation of the diode
capacitance with respect to reverse diode voltage is given in Fig. 8.9(b).
8.4. SUPERHETERODYNE RECEIVER 275
The circuit in Fig. 8.9(c) delineates the way a varactor diode is used in
a variable frequency oscillator circuit. The potentiometer Rtune , connected
between two supply voltages, control the reverse voltage bias on the diode. The
capacitance Co is a DC block capacitor, preventing a short-circuit of the DC bias
voltage by the inductance L. The large series resistor only serves to isolate the
tuned circuit elements from Rtune , so that the Q of the resonant circuit remains
high. The series combination of diode capacitance and Co appears across the
tank circuit. The total capacitance of the tank circuit becomes
CD Co
C+ (8.29)
CD + Co
Adjusting the potentiometer can now vary the resonance frequency of the tank
circuit. Since the oscillation frequency is determined by a voltage, this circuit
is called a voltage controlled oscillator (VCO).
Mixer
vRF(t) vIF(t)
BPF IF Amp
RF In
vLO(t)
LO
pose the input RF signal is a sinusoidal signal vRF (t) = VRF cos(ωRF t). A vari-
able frequency signal source is the local oscillator (LO). It generates a sinusoidal
signal vLO = VLO cos(ωLO t). The function of the mixer is to multiply these two
signals. The output of the mixer, vIF (t) is
vIF (t) = vRF (t) × vLO (t) = VRF VLO cos(ωRF t) cos(ωLO t) (8.30)
Table 8.1: IF, LO, RF and image frequencies of some commonly used systems
will be selected and be amplified, while all other signals will be rejected by the
BPF. As an example, let us choose fLO =12 MHz and fIF =15 MHz. In this
case, fRF =27 MHz is the selected frequency.
We note that there is one more RF frequency that can be selected:
since fRF i + fLO = fIF is the frequency of the first sinusoid of Eq. 8.31. This
frequency is called the image frequency. In the example above, the image fre-
quency is at fRF i =3 MHz. To prevent two distinct frequencies to be amplified,
we need to get rid of the image frequency at the RF input before it gets into the
mixer. For this purpose, we use a band-pass-filter to reject the image frequency.
Since the image frequency, fRF i , is far away from the desired fRF , the rejection
can be easily achieved using a simple BPF and it does not have to be a variable
frequency. In the example above, we need to reject 3 MHz while passing 27 MHz
signal.
Mixer
fRF fIF
RF In BPF BPF IF Amp
fLO
To reject image
frequency LO
Table 8.4 lists the IF, LO, RF and image frequencies for some commonly
used systems.
8.5. EXAMPLES 277
8.5 Examples
Example 4
We have a superheterodyne receiver with image rejection as in Fig. 8.11. fRF =
28 MHz, fLO = 27.5 MHz and fIF = 500 kHz. The BPF to reject the image
frequency is depicted in Fig. 8.12(a). The transformer is wound on a core
with AL = 2.1 nH/T2 . The antenna is represented with purely resistive source
impedance of 70 Ω. Find the value of Cs to receive at 28 MHz. What is the
image frequency? What is the image rejection in dB?
LO LO
(a) (b)
Figure 8.12: (a) Image reject filter of a superheterodyne receiver, (b) Compo-
nents transferred to the secondary side.
Solution
The source resistor of 70 Ω and the input voltage source can be transferred to the
secondary side as 70(10/2)2 = 1750 Ω and (10/2)Vin as depicted in Fig. 8.12(b).
Hence a maximum power transfer is achieved since the source resistor is equal
to the load resistor. The inductance of the secondary is Ls = 2.1 · 102 = 210 nH.
We can find Cs to resonate with Ls at 28 MHz using
25330
Cs = = 154 pF
282 · 0.21
The resulting circuit is a band-pass-filter with n = 1. Comparing with the LPF
prototype, we find from Eq. 6.17 in page 226
b1 2
∆f = = = 1.18 MHz
2πRCs 2π1750 · 154 · 10−12
The image frequency is at fRF i = fLO − fIF = 27.5 − 0.5 = 27 MHz. From
Eq. 6.24 of page 228 with n = 1, fo = 28 MHz and f = 27 MHz, we find
PL 1 1
= 2 2
= = 0.25
PA 1 + (fo /∆f ) (f /fo − fo /f ) 1 + (28/1.18) (27/28 − 28/27)2
2
Hence the rejection of the image frequency is 10 log10 (0.25) = −6.0 dB. Note
that it is not good idea to use the asymptotic approximations here, since the
frequency f is very close to the center frequency, fo and f < 0.1fo is not satisfied.
Example 5
Repeat the problem above when fLO = 25 MHz and fIF = 3 MHz.
8.5. EXAMPLES 278
Solution
In this case, the image frequency is at fRF i = fLO − fIF = 25 − 3 = 22 MHz.
We find the image rejection as
PL 1 1
= = = 0.0074
PA 1 + (fo /∆f )2 (f /fo − fo /f )2 1 + (28/1.18)2 (22/28 − 28/22)2
The rejection of the image frequency is 10 log10 (0.0074) = −21.3 dB, obviously
a better rejection value than the value found in the previous example.
Example 6
The OPAMP circuit shown in Fig. 8.13(a) acts like a square-wave oscillator.
Unlike the negative resistance oscillator of Fig. 8.4(b), the OPAMP operates
in the saturated output region. What is the frequency of the square wave, if
V + = V −?
R2 vo
+
V
R1
+
+
V vp
v1 vo T2
T1
v2 − v2
V - vn
C v1
v1 -
V
R
Solution
Referring to Fig. 8.13(b), at t = 0+ , the output voltage is vo = V + . Therefore,
the voltage, v1 (0+ ), at the positive input terminal of OPAMP is found from the
voltage divider as
R1
v1 (0+ ) = V + = vp
R1 + R2
At t = 0+ , the voltage, v2 , (equal to the capacitor voltage) at the negative input
terminal is assumed to be equal to an unknown voltage v2 (0+ ) = vn < 0 which
will be determined later.
The capacitor C charges toward V + with a time constant τ = RC. During
the charging period (0 < t < T1 ) we have v1 (t) > v2 (t), and hence the output
voltage remains at vo (t) = V + for 0 < t < T1 . Using the procedure of page 46,
we write
v2 (t) = V + + (vn − V + )e−t/τ for t > 0
When v2 (t) exceeds the voltage vp at t = T1+ , we get v2 (T1+ ) > v1 (T1+ ) and the
output voltage jumps to the negative saturation voltage V − . Thus the voltage,
8.5. EXAMPLES 279
At t = T2+ , v2 goes below vn and as a result the output voltage jumps back to
V + , completing the full cycle. We can find T1 using the exponential equations
above:
+
V − vn
v2 (T1 ) = vp = V + + (vn − V + )e−T1 /τ or T1 = τ ln
V + − vp
In a similar manner, we find T2 :
V − − vp
v2 (T2 ) = vn = V − + (vp − V − )e−(T2 −T1 )/τ or T2 − T1 = τ ln
V − − vn
Example 7
The OPAMP circuit depicted in Fig. 8.14 is known as phase shift oscillator.
Find the frequency of oscillation and the conditions for oscillation.
Solution
Assuming that the OPAMP is not saturated, we have V1 = V2 = 0. We write
the node equations for the nodes V3 , V4 and V5 as
V3 − Vo V3 V3 − V4
+ + =0 (8.34)
1/jωC R 1/jωC
V4 − V3 V4 V4 − V5
+ + =0 (8.35)
1/jωC R 1/jωC
and
V5 − V4 V5 1 + jωRC
+ =0 or V4 = V5 (8.36)
1/jωC R jωRC
8.5. EXAMPLES 280
V5 V4 V3
C C C
R R R
V2 − Rf
V1 Vo
+
To simplify the notation, we set X = jωRC. Combining Eqs. 8.34 and 8.36, we
get
X 1+X
V3 = Vo + V5 (8.37)
1 + 2X 1 + 2X
Combining Eqs. 8.35 and 8.36, we reach at
1 + 2X 1 + X
V3 = Vo − Vo (8.38)
X X
Equating the right hand sides of Eqs. 8.37 and 8.38
X 1+X (1 + 2X)(1 + X) − X 2
Vo + V5 = V5 (8.39)
1 + 2X 1 + 2X X2
After simplification we get
V5 X3 −jω 3 R3 C 3
= = (8.40)
Vo 1 + 5X + 6X 2 + X 3 (1 − 6ω 2 R2 C 2 ) + j(5ωRC − ω 3 R3 C 3 )
Rf ω 2 R 2 C 2
=1 (8.44)
R 5 − ω 2 R2 C 2
8.5. EXAMPLES 281
8.6 Problems
1. Consider a superheterodyne receiver with fIF =1 MHz and fRF =28 MHz.
(a) Find the image frequency. (b) Design an input band-pass-filter (as
shown in Fig. 8.11) that passes 28 MHz with an attenuation no more than
1 dB and rejects the image signal by at least 40 dB. (Assume that source
and load impedance of the filter is 50 Ω.)
2. Consider the circuit in Fig. 8.15(a). Find the transfer function H(ω) =
V2 (ω)/V1 (ω). Determine the frequency, ωo , at which ∠H(ωo ) = 0, when
R3 = R4 and C3 = C4 . What is |H(ωo )| at that frequency?
3. The circuit shown in Fig. 8.15(b) is called the Wien-bridge oscillator. Max
Wien (1866–1938) invented the Wien bridge. William Hewlett (1913–
2001) of Hewlett-Packard company was the first to build an Wien-bridge
oscillator as the first product of the company. The network shown in
Fig. 8.15(a) provides the positive feedback. The circuit oscillates at the
frequency ωo where ∠H(ωo )=0 (as given in problem 2), if the condition
(1 + R2 /R1 )|H(ωo )| > 1 is satisfied. If R3 = R4 and C3 = C4 , what is the
minimum value of R2 /R1 so that the circuit can oscillate?
4. Consider a parallel LC circuit with L=320 nH and a resonance frequency
of 28 MHz. A varactor diode is placed in parallel with it. When the
reverse voltage across the varactor diode is changed between 1 to 10 V,
the capacitance of the diode varies between 4 pF to 1.5 pF. Find the new
resonance frequency with the varactor reverse bias at 1 V and at 10 V.
5. A transconductance amplifier with Gf =0.02 S is connected as a Colpitts
oscillator as shown in Fig. 8.16(a). Find the value of the inductance if an
oscillation at 12 MHz is desired, and if C1 =10 pF and C2 =15 pF. Find
the maximum value of the resistance, r, that still allows an oscillation.
6. Repeat the problem 5, if there is a stray capacitance of Cs =7 pF exists as
shown in Fig. 8.16(b).
8.6. PROBLEMS 283
Gf Gf
L L
r C1 C2 r Cs C1 C2
(a) (b)
Figure 8.16: (a) Circuit for problem 5, (b) circuit for problem 6.
Chapter 9
ON THE AIR
Danish physicist Hans Christian Øersted (1777–1851) showed in 1820 that a cur-
rent carrying wire creates a magnetic field. The same year, André Marie Ampére
found that two parallel current-carrying wires can attract or repel each other,
depending on the relative direction of the currents. In 1831, Michael Faraday
observed that a moving magnet through a loop of wire can create a current.
Scottish physicist James Clerk Maxwell (1831–1879) was the first to formulate
the relations between electric and magnetic fields in a unified theory. His for-
mulation of 1865 shows that a changing magnetic field creates an electric field,
a changing electric field creates a magnetic field and predicts that the combina-
tion of these fields, called electromagnetic waves, propagate at the speed of light.
Four elegant vector equations* describing this behavior is known as Maxwell’s
equations, which underpins much of the modern technological world. In 1888,
Heinrich Rudolf Hertz confirmed experimentally the existence of electromag-
netic waves predicted by Maxwell.
Antennas convert electrical signals into electromagnetic waves for transmis-
sion, and they also work in the other direction to convert electromagnetic waves
to electrical signals for reception. Antennas are combinations of pieces of con-
ductors of specific lengths and shapes. There are many different types of anten-
nas for numerous applications. More detailed information on antennas can be
found in other books [15–17].
What follows in this chapter is a descriptive theory of electromagnetics and
antennas.
When the circuit dimensions are small compared to the wavelength, most
of the electromagnetic energy generated by the circuit is confined to the cir-
cuit. It is either conserved for the desired purpose or converted to heat. When
the dimensions of the components or the interconnections become large (e.g.,
comparable to the wavelength) part of the energy escapes into space in form of
electromagnetic waves. This part of the energy used in the circuit appears as
lost energy to the circuit, whereas it provides a source for the electromagnetic
waves in space.
Antennas are devices, which makes use of this conversion-and-escape mech-
anism to produce radio waves as efficiently as possible.
across the gap, as shown in Fig. 9.1, an electric field between the entire surfaces
of two conductors is produced due to the potential difference between them.
This field is time varying at the frequency of the source. The electric field is
denoted by letter E and it has units of V/m.
The electric field extends to the entire space, but its strength decreases as
the observation distance from the dipole increases. The electric field is strongest
near the gap.
The two conductor surfaces constitute a distributed capacitance across the
terminals at which the voltage source is applied. This phenomenon is depicted
in Fig. 9.2(a). The current Is supplied by the source, leaks from one conductor
to the other along the length of the conductor, through the capacitive path.
The current amplitude decreases as we move along the conductor. Current
diminishes at the tip of the conductor. A typical current distribution along
9.1. ANTENNA CONCEPT 286
the antenna is given in Fig. 9.2(b). The current along the antenna produces
Figure 9.2: (a) Distributed capacitance on antenna, and (b) current amplitude
distribution along antenna.
a magnetic field shown in Fig. 9.1. Again, the frequency of the magnetic field
is the same as that of the voltage source. Letter H denotes the magnetic field
and it has units of A/m. The electric and magnetic fields have magnitude and
direction and hence they can be modelled as vectors. Fig. 9.3 depicts color-
coded E and H fields in the vicinity of a dipole antenna as determined from a
finite-element simulation .
Figure 9.3: Magnitude of E-field (left) and y-component of H-field (right) shown
in z-x plane for a dipole antenna placed along z-axis (along horizontal direction).
Figure 9.4: Color coded E-field plots of a 28 MHz dipole antenna (5.35 m long)
placed along z-axis (vertical direction). E-fields are shown at planes parallel to
z-axis at distances 0.25, 0.5, 1.0, 2.0, 4.0 and 8.0 m away.
centric spheres. The center of these spheres, which is called the phase center,
coincides with the center of the isolating gap in the dipole. This is shown in
Fig. 9.5. The direction of propagation in this figure is outward from the center.
Figure 9.5: Radio wave far away from the source dipole.
The electromagnetic wave generated at some instant gets away from the dipole
in all directions, at a speed of 3·108 m/sec.
Now let us take a closer look at the field shown on the patch over the spherical
surface, in Fig. 9.5. If the radius of the sphere is very large compared to the
rectangular patch (which is a very realistic assumption for practical antenna
discussions), the patch approximately defines a planar surface. We can define a
Cartesian plane on which electric field coincides with x axis and magnetic field
coincides with y axis. This is shown in Fig. 9.6. With Ex and Hy are phasors
of the electric and magnetic fields respectively, and ax and ay are unit vectors
9.1. ANTENNA CONCEPT 288
Figure 9.6: Orthogonal electric and magnetic fields and the direction of propa-
gation.
where Po is the power delivered to the antenna and 4πr2 is the area of the
sphere. From Eq. 9.7 and 9.8 the electric field at that distance becomes
r √
p 2ηo Po 60Po
|Ex (r)| = 2ηo P (r) = = (9.9)
4πr2 r
where Po is in watts and r is in meters. For example, a transmitter delivering
10 mW to an omnidirectional (no directivity) antenna generates an electric field
strength of 0.8 mV/m at 1 km distance.
- Vr +
2a
2l
E H
Figure 9.7: A dipole of total length 2l correctly aligned to receive the incoming
electromagnetic field.
I2
Po = Rr (9.18)
2
For the same input current I, a dipole (of length 2l ) radiates a total power Pr
to the entire space, while a monopole (of length l) radiates only to half space,
hence a total power of Pr /2. Therefore, the radiation resistance of monopole is
half as much as the radiation resistance of an equivalent dipole. For example, a
monopole of length 10 cm at 28 MHz has a radiation resistance of 0.034 Ω. If
the length is 50 cm, the radiation resistance becomes 0.86 Ω.
Similarly a monopole antenna of length λ/4 placed on a large conducting
surface behaves like a λ/2 long dipole.
The effective capacitance between the conductor (of length l and radius a)
and the infinitely large ground plane in a monopole antenna is given by
l l
Cm ≈ 2πϵo for < 0.1 (9.19)
ln(l/a) − 1 λ
where Cm is in Farads and both l and a are in meters. Therefore, the capacitance
of a monopole of length l is the twice as much as the capacitance of a dipole of
length 2l. For example, the capacitance of a monopole of length 10 cm with a
diameter of 1 mm is 1.29 pF.
When the monopole is not short, half of the reactance given in Eq. 9.14 can
be used in the equivalent circuit of the monopole:
Xd l XL (l/λ) l
Xm = = −60 cot 2π [ln (l/λ) − 1] + for < 0.25 (9.20)
2 λ 2 λ
Fig. 9.9 shows the series resistance, Rm , and series reactance, Xm , of a monopole
antenna using Eqs. 9.17 and 9.20 for various l/a ratios. For example, a monopole
9.2. DIPOLE ANTENNA IMPEDANCE 292
−300 25
−600 20
Rrm (Ω)
Xm (Ω)
−900 15
−1200 10
−1500 5
−1800 0
0 0.05 0.1 0.15 0.2 0.25
l/λ
Figure 9.9: Series resistance (solid) and series reactance with l/a=100 (dash-
dot), l/a=1000 (dotted), l/a=10000 (dashed) of a monopole antenna placed on
a very large planar conducting ground plane.
Rout Rrm
+ Rrm Xm
Rin
+
Xm Vr
(a) (b)
In fixed stations where the antennas are installed in places like roofs of
buildings, it is possible to simulate a ground plane to an acceptable level by
properly designed conducting frame. It is almost never possible to have a large
planar conducting surface (compared to the wavelength), on which an antenna
can conveniently be placed in mobile stations. The radiation resistance of a
monopole is always determined by the mutual impedance of the ground reference
in such systems. This can be limited to the dimensions of the casing of a handset
in case of mobile phones.
Ionosphere
300km
Earth
longer at night when the lower layers of ionosphere disappear. This is why AM
radios receive higher number of stations at night.
Electromagnetic waves at frequencies between 3 MHz and 30 MHz (wave-
lengths between 100 m to 10 m), are reflected by the higher layers of ionosphere
(up to 500 km high) if their frequencies are below a critical frequency, fc . The
critical frequency is determined by the ionosphere conditions and hence it de-
pends on the time of the day, the season of the year and the year of 11-year
cycle. fc is typically less than 10 MHz, but it may go as high as 30 MHz. Higher
frequencies are attenuated less at the reflections of the imperfect conductors of
the ionosphere and the earth surface. Therefore, the best propagation occurs
at frequencies just below the critical frequency. For example, if fc > 28 MHz,
a communication at 28 MHz may be possible with a small power (1 W) trans-
mitter at distances as high as 4000 km.
Above the critical frequency, fc , no reflection at the ionosphere occurs and
the electromagnetic waves penetrate through it. Hence such electromagnetic
waves (for example, FM radio band) can be received only in line-of-sight mode.
However, during the hot days of summer, especially in Mediterranean Sea and
the Persian Gulf regions, it may be possible to receive FM signals from distances
1000 to 4000 km due to a type of radio propagation known as tropospheric
ducting. In this mode of propagation, the waves do not travel in straight lines
but rather in curves, due to an abnormal distribution of temperature in the
atmosphere.
requirements of this course, you must contact your national regulatory agency
and obtain an amateur license.
9.4 Problems
1. A short monopole of length 20 cm is used as the antenna of a mobile set
operating at 150 MHz. What is the open circuit voltage generated at a
receiving antenna, when another set transmits 2W at 3km distance?
LTSpice Tutorial
It is a very powerful program and can be used to simulate any linear or nonlinear
circuit of any size. It is highly recommended that you learn how to use it
effectively.
The schematic entry can use more than 2,000 symbols. You can also draw
your own symbols for devices you wish to import into the program.
Transient (Time domain) analysis can be used to simulate circuits containing
linear (R, L or C) or nonlinear (diodes) elements of any order. Note that the
analytic analysis methods presented in the text deals only with time-domain
analysis of first-order circuits. LTSpice can cope with any order.
AC analysis can be used to find the transfer function of circuits containing
linear elements only. LTSpice solves the circuits using the phasor approach.
Press F2 (or click Component icon) → voltage (to enter a voltage source),
left-click to enter it on the schematic window.
Right click on the voltage source → 5 (to set the voltage at 5 V)
Right click on the resistor → 2 kΩ (to define the value of resistor as 2 kΩ)
Figure A.1: LTSpice schematic of a resistive circuit and the result window after
DC analysis.
Ctrl-R twice to flip the current source before placing it in the schematic.
Right click on the current source → 4 m (to set the current at 5 mA)
Press g (or click Ground icon) → place it on the bottom node (to define
the ground node)
Press F3 (or click Wire icon) (to join the components using Wire tool)
Press F4 (or click Label Net) → Va and place it above V1 as the node
name. Two nodes with the same node name are assumed to be connected
(without a wire between them). This is useful to simplify the schematic if
it becomes too crowded with wires.
A result window will appear as in Fig. A.1 showing the voltage and current
values.
The reference direction of currents of resistors are determined by the orig-
inal placement or rotation of the component. If the reference label of a
resistor is on the right-hand-side, the current is defined downwards, oth-
erwise it is upwards. To modify the reference direction click Move icon,
choose the resistor by placing a rectangle over it and ctrl-R to rotate it.
A.2. INPUT-OUTPUT RELATION OF A DIODE CIRCUIT 299
Place the Ground node and join the components using Wire tool.
Press F4 (or click Label Net icon) → Va (to label the input node)
An empty graph should appear. You can add a trace by going to schematic
window and clicking on the node when red voltage probe appears. Click
Vb node to see the relation Vb versus Va .
Plot Settings → Notes and Annotations → Place Text: Va. Place it near
X-axis.
You can copy the graph using Tools → Copy bitmap to Clipboard (for
pasting into another application)
Figure A.4: LTSpice simulation result of capacitor voltage for the RC circuit.
You can attach cursors to read the values of the graphs. For this purpose,
right click on the trace name and click Attached cursor combo box. You
can attach one or two cursors. Using two cursors, you can determine the
difference between two points on the graph.
In a time domain simulation, the instantaneous power dissipated on a
component can be determined by clicking ALT-Left click on the compo-
nent. An expression for the instantaneous power (involving voltage and
current) will display in the graph window. The average value of this ex-
pression can be found by CTRL-Left click on the instantaneous power
expression in the graph window. Note that for an accurate average value
the simulation time should be an integer multiple of the excitation period.
When a lossless component like a capacitor is in a circuit with sinusoidal
excitation, the instantaneous power on the capacitor in the steady-state
will change between positive and negative values with an average value of
zero. (Because of numerical errors, that average value might be a small
but a non-zero value.) If the circuit is not excited sinusoidally, the power
may be non-zero, indicating the finite stored energy in the capacitor.
You can change the colors of the graphs using Tools → Color Preferences
Change the initial condition of the capacitor and the value of the resistor.
Find out what happens.
You can also specify an initial condition on a branch current. For example,
IR1(0)=1 mA can be specified by .IC i(R1)=1 m
Press Run command. Click on the Vin, VC1 and VC2 nodes to display
the voltage waveforms as shown in Fig. A.6. Note that the waveforms
are not of the kind that are solved analytically in the text. LTSpice can
handle circuits of any order :)
Figure A.6: LTSpice simulation of the time domain response to a pulse for the
second-order RC circuit.
Press Run command. You should have the graph as shown in Fig. A.8.
The left Y-axis shows the magnitude in dB, while the right Y-axis shows
Figure A.8: LTSpice simulation of the transfer function amplitude (solid) and
phase (dotted) of the RLC circuit.
Copy the plot using Tools → Copy bitmap to Clipboard to paste it into
another application.
R3
R2
V3 10K V1
AC 1 1K U1
R1
vout 12
1K LT1413
V2
12
Draw the schematic shown in Fig. A.9. To choose the OPAMP, click on
Component icon, choose [Opamps]. Then click on LT1413. Note that
LTSpice contains the models for components produced by Analog Devices
only. You can add other models if SPICE models are available.
Write Start (1K for 1 KHz) and Stop (100Meg for 100 MHz) frequencies.
Place the .ac dec 100 1Meg 30Meg command at the schematic.
Press Run command. You should have the graph as shown in Fig. A.10.
The gain at low frequencies is 20 dB as expected. But the gain drops
above 100 KHz, because of the OPAMP limitations. The phase shift at
low frequencies is nearly 180o , indicating the inversion.
To perform a large signal analysis, enter a large signal excitation to the
input source by Right-Click: SINE, Amplitude[V]: 1.5, Freq[Hz]: 1K
A.7. ANALYSIS OF A BJT CIRCUIT 305
V(vout)
24dB 180°
18dB 160°
12dB 140°
6dB 120°
0dB 100°
-6dB 80°
-12dB 60°
-18dB 40°
-24dB 20°
-30dB 0°
-36dB -20°
-42dB -40°
-48dB -60°
1KHz 10KHz 100KHz 1MHz 10MHz 100MHz
Figure A.10: LTSpice simulation of the transfer function amplitude (solid) and
phase (dotted) of the inverting amplifier circuit.
R3
R2
V3 10K V1
AC 1 1K U1
R1
vout 12
1K LT1413
SINE(0 1.5 1k) V2
.tran 4m 12
Figure A.11: LTSpice schematic of the OPAMP circuit for transient analysis.
In Edit Simulation Command, click Transient analysis tab. Set the Stop
time to 4m, to see four cycles. The schematic window should look like in
Fig. A.11.
Press Run command. You should now have the graph showing the clipped
sine wave as shown in Fig. A.12.
First perform a transient analysis to make sure that the BJT operates
in ACT region. For the input source, choose Sine wave in the Advanced
tab. Choose the frequency as 1K (1 KHz) and the peak amplitude as 10m
(10 mV). Perform a transient analysis for 10m (10 ms). Observe that the
A.7. ANALYSIS OF A BJT CIRCUIT 306
V(vout)
12V
10V
8V
6V
4V
2V
0V
-2V
-4V
-6V
-8V
-10V
-12V
0.0ms 0.4ms 0.8ms 1.2ms 1.6ms 2.0ms 2.4ms 2.8ms 3.2ms 3.6ms 4.0ms
Figure A.13: LTSpice schematic of the BJT amplifier circuit for transient and
AC analysis.
Figure A.14: LTSpice simulation of the transient response of the BJT amplifier.
Figure A.15: LTSpice simulation of the BJT amplifier with a larger input signal.
Figure A.16: LTSpice simulation of the gain amplitude (solid) and phase (dot-
ted) of the BJT amplifier circuit.
Appendix B
Significant figures
The rules for determining the number of significant figures in a given problem
can be stated as
1. For addition and subtraction of numbers with known significant figures,
the result should have as many decimal places as the number with the
smallest number of decimal places. For example, 145.2 + 1.245 = 146.4 or
4.78 · 10−2 − 2.678 · 10−4 = 4.75 · 10−2
2. For multiplication and division of numbers with known significant figures,
the results should have as many significant figures as the number with the
smallest number of significant figures. For example, 20.11 √ × 1.2 = 24 or
1.854 · 10−8 /5.35 · 104 = 3.47 · 10−13 . Hence,
√ writing 10.1 210.1 = 14.284
is poor engineering. One should write 10.1 2 = 14.3.
Unit prefixes
Symbol Name Magnitude Symbol Name Magnitude
y yocto 10−24 da deca 101
z zepto 10−21 h hecto 102
a atto 10−18 k kilo 103
f femto 10−15 M mega 106
p pico 10−12 G giga 109
n nano 10−9 T tera 1012
µ micro 10−6 P peta 1015
m milli 10−3 E exa 1018
c centi 10−2 Z zetta 1021
d deci 10−1 Y yotta 1024
Appendix C
Answers to selected
problems
Chapter 1
1. λ=17 m and λ=1.7 cm
3. 34 dB.
4. 2.0 cos(ωo + ωs )t + 2 cos(ωo − ωs )t
5. Mixer output:
1
(cos(ωo − ωs )t + cos(ωo + ωs )t) + cos(ωo − 2ωs )t + cos(ωo + 2ωs )t
2
3
+ (cos(ωo − 3ωs )t + cos(ωo + 3ωs )t) + 2(cos(ωo − 4ωs )t + cos(ωo + 4ωs )t)
2
Chapter 2
1. (a) 5.6 k±10%, (b) 390 k±10%, (c) 750 Ω±5%
2. PR =0.27 W, PV =0.27 W, PI =0.
√
3. 5/ 3.
4. (a) 46 Ω (b) 42 Ω (c) 8.4 k (d) 2.0 k (e) 49 Ω
5. (a) 7.1 (b) 0.99 (c) 20.
6. (a) 159 Hz (b) 50.0 Hz (d) ω/2π
10. (a) I1 =10.9 mA, I2 =1.07 mA, V1 =54.6 mV, (b) I1 =2.96 mA, I2 =0.926 mA,
V1 =2.04 V, (c) I1 =96.5 mA, I2 =1.66 mA, V1 =12.6 V, (d) I1 =-1.29 mA,
V1 =-4.27 V, (e) I1 =-0.467 mA, I2 =-7.85 mA, V1 =-1.54 V
11. (
−89.2 + 119.2e−t/21.3 for 0 < t < 40 µs
iL (t)(mA) =
0.8 − 71.8e−(t−40)/21.3 for t ≥ 40 µs
(
−12 − 111.7e−t/21.3 for 0 < t < 40 µs
vR (t) = −(t−40)/1.33
−12 + 1077e for t ≥ 40 µs
12. (
0.1 − 0.095e−t/416 for 0 < t < 100 µs
iL (t)(mA) =
−0.1 + 0.1253e−(t−100)/416 for t ≥ 100 µs
19. (a) iL (0− ) = −2.0 mA, vL (0− ) =0.0, (b) iL (0+ ) = −2.0 mA, vL (0+ ) =3.3 V,
(c) vL (t) = 0.0 for t < 0, and vL (t) = 3.3e−t/4.5µ for t ≥ 0.
20. Voltage gain=20. dB. When Rin = RL , power gain=100=20. dB. When
Rin = 10RL , power gain=1000=30. dB.
Chapter 3
1. (a) -59.1-j9.39=59.9∠(−171o ), (b) -2.43+j3.11=3.94∠128o ,
(c) 0.630+j3.92=3.97∠80.9o ,
(d) 2.54·10−3 − j4.58 · 10−4 = 2.58 · 10−3 ∠(−10.2o )
8. (a) 140∠ − 0.79, (b) 33∠ − 0.79, (c) f = 0: 94; f = 100 kHz: 93∠ − 0.069;
f = 500 kHz: 82∠ − 0.27; f = 1 MHz: 67∠ − 0.34; f = 1.44 MHz: 59∠ −
0.32; f = 5 MHz: 48∠ − 0.14; (d) 140∠0.78, (e) f = 1 MHz: 150∠0.020;
f = 19 MHz: 180∠0.16; (f) f = 1 MHz: 12∠1.6; f = 36.5 MHz: 390∠1.4;
f = 73 MHz: 720∠1.5.
9. For V =5∠π/6 V: (a) I=0.036∠1.3 ⇒ i(t) = 0.036 cos(2π7.2 · 106 t + 1.3).
(b) I=0.15∠1.3 ⇒ i(t) = 0.15 cos(2π720 · 103 t + 1.3). (d) I=0.036∠ − 0.26
⇒ i(t) = 0.036 cos(2π7.2 · 106 t − 0.26). (e) f = 1 MHz: I=0.0333∠28.9o
⇒ i(t) = 0.0333 cos(2π106 t + 28.9o ).
Chapter 4
1. (
2.3 − 2.3e−t/0.002 for 0 < t ≤ 0.0020
vC (t) =
1.45 for t ≥ 0.0020
3. From the voltage doubler circuit of Fig. 4.40 in p. 172, we have VC1 = 9.4 V
and VC2 = 18.8 V. Hence v1 = 10 sin(ωt) − 9.4. C3 and D3 forms a
voltage clamper for this input ⇒ VC3 =18.8 V. D4 and C4 forms a half-wave
rectifier ⇒ VC4 = 18.8 V. The output voltage is vo = VC2 + VC4 =37.6 V.
This circuit is a voltage quadrupler.
4. (a) ACT, 6.92 V, (b) ACT, 6.72 V, (c) ACT, 6.61 V.
8. SAT, −4.28 V.
9. ACT, 6.96 V.
APPENDIX C. ANSWERS TO SELECTED PROBLEMS 315
Chapter 5
2. VR = 5.94∠53.5o , VL = 104∠143.5o , VC = 112.5∠ − 36.5.
3. VR = 10∠0, VL = 51∠90o , VC = 51∠ − 90o .
4. C=98.9 pF
5. ωo = 95.35 · 106 rps ω1 = 85.87 · 106 rps ω2 = 105.87 · 106 rps
7. " 2 #1/2
1 RL
ωo = −
LC L
p
If RS > L/C, no resonance occurs.
9. Q=10.3
10. R=150 Ω, C=211 pF.
11. C=2.27 nF. 176 mW.
12. For R=0 fo =15.915 MHz. For R=1 fo =15.914 MHz. For R=100 no
resonance.
13. fo =1.33 MHz.
14. C=24.8 pF. vR (t) = 5 cos ωo t, vL (t) = 11.4 cos(ωo t + 90o ),
vC (t) = 11.4 cos(ωo t − 90o ).
15. Lower 3-dB: f1 =22.6 MHz, vR (t) = 3.5 cos(ω1 t−45o ), vL (t) = 6.5 cos(ω1 t+
45o ), vC (t) = 10.0 cos(ω1 t − 135o ).
Upper 3-dB: f1 =34.8 MHz, vR (t) = 3.5 cos(ω1 t + 45o ),
vL (t) = 10.0 cos(ω1 t + 135o ), vC (t) = 6.5 cos(ω1 t − 45o ).
16. RDC =54 mΩ. fsc =2 MHz. R=200 mΩ at 28 MHz.
17. l=9.8 mm.
18. N =9.
19. 18 turns
20. AL =2.65 nH/T2 . R=0.24 Ω.
23. At f =400 kHz, Vout /Vin =9.6∠ − 29o .
At f =1 kHz, Vout /Vin =11∠ − 0.08o .
24. At f =1 MHz, Vout =4.08∠ − 172o .
At f =1 kHz, Vout =4.4∠ − 150o .
25. C=99 pF |B1 /A1 | = 0.09 |B2 /A2 | = 0.007
APPENDIX C. ANSWERS TO SELECTED PROBLEMS 316
Chapter 6
1. Vout =1.52∠ − 0.1o , IC = 12∠89.9o
2. C1 = C3 = 53 nF L2 = 0.1 H
7. C1 =2.1 nF L1 =13.3 nH
8. 75 Ω.
9. Z=48.2∠50o 50o
Chapter 7
1. C < 0.23 µF.
Chapter 8
2. H(ω) = jωR3 C3 /(1 − ω 2 R32 C32 + 3jωR3 C3 ), ωo = 1/R3 C3 , |H(ωo )| = 1/3
3. R2 /R3 > 2.
Chapter 9
1. 0.73 mV
2. 2l=209 cm, Cd =6.66 pF. L=5.22 µH. Q=126.
Bibliography
[4] R. White and R. Doering, Electrical Engineering Uncovered, 2nd ed. Pear-
son, 2001.
[5] N. Balabanian, Electric Circuits. McGraw-Hill, 1995.
[6] S. Sze and M. Lee, Semiconductor Devices: Physics and Technology, 3rd ed.
John Wiley & Sons, Inc., 2012.
[7] R. Gayakwad, Op-Amps and linear integrated circuits, 4th ed. Prentice-
Hall, 1999.
[8] K. Blair, Audio Engineering Handbook. McGraw-Hill, 1988.