0% found this document useful (0 votes)
14 views87 pages

Lecture Material 2023

The document provides an overview of operational amplifiers (op-amps), detailing their properties, ideal characteristics, and practical applications in electronic circuits. It explains the basic configurations of op-amps, including non-inverting, inverting, and differential amplifiers, along with their voltage gain and input/output relationships. Additionally, it highlights the significance of op-amps in various electronic functions and their role as fundamental building blocks in analog computing.

Uploaded by

desmond4252
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
14 views87 pages

Lecture Material 2023

The document provides an overview of operational amplifiers (op-amps), detailing their properties, ideal characteristics, and practical applications in electronic circuits. It explains the basic configurations of op-amps, including non-inverting, inverting, and differential amplifiers, along with their voltage gain and input/output relationships. Additionally, it highlights the significance of op-amps in various electronic functions and their role as fundamental building blocks in analog computing.

Uploaded by

desmond4252
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 87

CHAPTER 1

THE OPERATIONAL AMPLIFIER

1.1 Properties of Operational Amplifier

An Operational Amplifier (OP-AMP) is a multi-stage amplifier with high gain (typically 200,000),
in which feedback is added to control its overall response characteristics. It consists of a complex
arrangement of resistors, transistors, capacitors and diodes, and is treated as a single device. The
Op-Amp is used to perform several different functions and it forms the basic building block of
many electronic circuit applications. The name “operational amplifier” originates from the use of
this type of amplifier to perform specific electronic circuit functions or mathematical operations
of addition, subtraction, multiplication, division, differentiation, and integration, by using voltage
as an analogue of another quantity. This is the basis of the analogue computer where op-amps
were used to model the basic mathematical operations.

An operational amplifier can amplify signals having frequency ranging from 0 𝐻𝑧 to about
1 𝑀𝐻𝑧. This means that op-amps can be used to amplify dc as well as ac input signals.
Generally, they are thought of as universal “gain blocks” whose function within a circuit could be
defined with the addition of external components. For example, with the addition of two resistors,
an amplifier with a defined gain can be produced. Likewise, a simple low-pass filter could be
produced with a single resistor and a capacitor.

Presently, the Op-Amp is the most versatile and widely used of all linear integrated circuits; it is
now more of a circuit building block, and uses relatively low supply voltages. For example, the
OP-Amp can now be a complete amplifier circuit constructed as an IC on a single silicon chip.
There are other packages that are available commercially in several forms. Inside these packages
are a number of transistors and other components integrated to form a unit. A typical one is the
eight-pin dual in-line package (or DIP), which is depicted in Figure 1.1:

Balance 1 8 No connection

Inverting input 2 7
V+

Non-inverting input 3 6 Output

V- 4 5 Balance

Figure 1.1: A Typical Op-Amp Pin Configuration

Op-amps are also used in phase shifting, voltage regulation, analog computer operations, in
instrumentation and control systems, oscillator circuits, pulse generators, square-wave generators,
triangular-wave generators, comparators, analog-to-digital and digital-to-analog converters (ADC
and DAC); voltage-to-current converters, current-to-voltage converters, sample-and-hold circuits
etc.
A typical op-amp symbol as shown in Figure 1.2 has five important terminals. These are:

1
i. The inverting input, pin 2
ii. The non-inverting input, pin 3
iii. The negative power supply V-, pin 4
iv. The output, pin 6
v. The positive power supply V+, pin 7

V+
7

Inverting input 2 -
6 Output
Non-inverting input 3 +

4 1 5
V-
Figure 1.2: A typical op-amp circuit symbol (741 op-amp IC)

From Figure 1.2, a few important points about an op-amp can be drawn:
1. The op-amp has two input terminals, the inverting input (-) and the non-inverting input
(+), and one output terminal.
2. The word ‘non-inverting’ means that if a signal is applied at this terminal of the op-amp
it will appear with the same polarity at the output (output in phase with the input signal).
3. An input applied to the inverting terminal will appear inverted at the output (output 1800
out of phase with the input signal).
4. The op-amp amplifies the difference between the voltages applied at the non-inverting
input and the inverting input. The difference between the two voltages acts as input to
the op-amp. This input difference could be denoted as 𝑉𝑖𝑛 .
5. The typical op-amp operates with two dc supply voltage, one positive and the other
negative. Usually, these dc voltage terminals are left off the schematic symbol for
simplicity but are always understood to be there.
In Figure 1.2, this op-amp is an eight-pin type, but terminals (pins) 1 and 5 are of little
importance to us at the moment, even though such pins could be used for frequency
compensation. Terminal 8, however, is unused.

1.2 The Ideal Op-Amp


An op-amp is said to be ideal if it possesses the following characteristics:
i. It has infinite gain (open loop voltage gain, 𝐴𝑜𝑙 = ∞). Infinite voltage gain means the
voltage difference (Vin) required between the two inputs to produce any output voltage is
zero.
ii. It has infinite input resistance (impedance). The input resistance is the resistance measured
between the inverting and non-inverting terminals (𝑅𝑖𝑛 = ∞).
iii. The output resistance seen looking back into output terminals is zero (impedance, Rout =
0), that is, output voltage, Vout is independent of the load resistance connected between the
output terminals.

2
iv. It should have infinite bandwidth (flat frequency response from dc to infinity Hz). Thus,
any frequency signal can be amplified without attenuation.
v. The voltage gain remains constant over a wide frequency range.
vi. Zero voltage output level for a zero input level, i.e. offset voltage is zero.

vii. It has infinite slew rate (the maximum rate of change of output voltage with respect to time
– specified in 𝑉/𝜇𝑠)
viii. Insensitivity to power supply voltage variations

Table 1.1 Characteristics of an Ideal Operational Amplifier

S/N Characteristics Value


1. Voltage Gain ∞
2. Input Resistance ∞
3. Output Resistance 0
4. Bandwidth ∞
5. Offset Voltage 0
6. Slew Rate ∞

V1 i1 -

Vin Rin = ∞ AVin Vo


Ro = 0
A=∞

V2 i2 +

Figure 1.3: An ideal operational amplifier

From Figure 1.3, 𝑉𝑖𝑛 is the differential input voltage, which is the difference between the voltage
between non-inverting terminal and ground (V2) and the voltage between inverting terminal and
ground (V1). That is
𝑉𝑖𝑛 = 𝑉2 − 𝑉1 (1.1)

The op-amp senses the difference between the two inputs, multiplies it by the gain A, and causes
the resulting voltage to appear at the output. Thus the output Vo is given by

𝑉𝑜 = 𝐴𝑉𝑖𝑛 (1.2)

Or
𝑉𝑜 = 𝐴(𝑉2 − 𝑉1 ) (1.3)

3
For circuit analysis, two important characteristics of the ideal op-amp are taken into consideration.
These are:

Assumption 1
The currents into both input terminals are zero
𝑖1 = 0; 𝑖2 = 0 (1.4)

This is as a result of the infinite input resistance.


That is,
𝑉 𝑉
𝑖= = =0
𝑅 ∞

This means that an open circuit exists between the terminals and current cannot flow into the op-
amp.
This does not necessarily mean the output current is zero (due to i+ and i- from V+ and V-
respectively).

Assumption 2
The voltage across the input terminals is negligibly small
𝑉𝑖𝑛 = 𝑉2 − 𝑉1 ≃ 0

⇒ 𝑉1 = 𝑉2 (1.5)

This assumption emanates from the fact that for the ideal op-amp, voltage gain, 𝐴 = ∞, and from
equation 1.2,
𝑉𝑜
𝑉𝑖𝑛 = , thus 𝑉𝑖𝑛 = 0
𝐴

In effect, the current into the two input terminals of an ideal op-amp is zero, and there is also
negligibly small voltage between the input terminals.

Although it is impossible to realise the ideal operational amplifier, its conceptual use allows us to
understand the basic performance to be expected from a given analog circuit and serves as a model
to help in circuit design. Once the properties of the ideal amplifier and its use in basic circuits are
understood, then various ideal assumptions can be removed in order to understand their effect on
circuit performance.

1.3 The Practical Operational Amplifier


Modern integrated circuit (IC) op-amps approach parameter values that can be treated as ideal in
many cases, but in practice no op-amp can be ideal. All devices have limitations and the IC op-
amp is no exception. Op-amps have both voltage and current limitations.
Thus characteristics of a practical op-amp are:
▪ High voltage gain (typical range = 10 5 to 107)
▪ High input resistance (typical range = 106 (typical bipolar op amp) to 1012Ω, (typical JFET
op amp).
▪ Low output resistance (10 to 100 Ω)

4
i1
V1 -

Ri AVin Vo
Vin n
Ro

V2
i2 +

Figure 1.4: A Practical operational amplifier

1.4 Open-Loop Voltage Gain (Aol)


The voltage gain, A of an amplifier describes the relation between changes in the input signal and
changes in the output signal. When an op-amp is operated without any connection between the
output and any of the inputs (i.e. without feedback), it is said to be in the open-loop condition.

-
Vi Vo
+

Figure 1.5: Open-loop operational amplifier


Open-loop voltage gain is the internal voltage gain of the op-amp measured in the differential
mode. This represents the ratio of output voltage to input voltage when there are no external
components. The open-loop voltage gain is set entirely by the internal design and can range all
the way to 200,000 or even more. This is not a well-controlled parameter. The open-loop voltage
gain is sometimes referred to as large signal voltage gain.
𝑉𝑜
𝐴𝑜𝑙 = (1.6)
𝑉𝑖𝑛

The following are the three open-loop operational amplifier configurations:


(a) non-inverting amplifier
(b) inverting amplifier
(c) differential amplifier

(a) Non-inverting
𝑅
𝑉2
+
𝑉𝑂
𝑉1 -
𝑣𝑖

Figure 1.6: Open-loop Non-inverting operational amplifier

5
In the non-inverting configuration as could be observed in the open-loop non-inverting amplifier
in figure 1.6, input is applied at the non-inverting terminal of the op-amp whilst the inverting
terminal is grounded.
𝑉𝑜
Now open-loop gain, 𝐴𝑜𝑙 = (from (1.6))
𝑉𝑖𝑛

Thus 𝑉𝑜 = 𝐴𝑜𝑙 𝑉𝑖𝑛 = 𝐴𝑜𝑙 (𝑉2 − 𝑉1 ) (1.7)


But from figure 1.6, 𝑉1 = 0
And if source resistance R is assumed very small, then it can be neglected
Thus 𝑉2 = 𝑣𝑖
⟹ 𝑉𝑜 = 𝐴𝑜𝑙 (𝑣𝑖 − 0)
Or 𝑉𝑜 = 𝐴𝑜𝑙 𝑣𝑖 (1.8)
Equation (1.8) shows that the output is A times larger than the input and in phase with the input.

(b) Inverting amplifier

𝑅
𝑉1
-
𝑉𝑂
+
𝑉2
𝑣𝑖

Figure 1.7: Open-loop Inverting Operational Amplifier

Figure (1.7) is an open-loop inverting amplifier. In inverting configuration, input is applied to


the inverting terminal of the op-amp whilst the non-inverting terminal is grounded.
Thus 𝑉2 = 0
But from equation (1.7), 𝑉𝑜 = 𝐴𝑜𝑙 𝑉𝑖𝑛 = 𝐴𝑜𝑙 (𝑉2 − 𝑉1 )
Thus if source resistance R is neglected, then
𝑉1 = 𝑣𝑖
⟹ 𝑉𝑜 = 𝐴𝑜𝑙 (0 − 𝑣𝑖 )
Or 𝑉𝑜 = −𝐴𝑜𝑙 𝑣𝑖 (1.9)

From equation (1.9), the output is A times larger than the input and out of phase with the input.

(c) Differential amplifier


In the case of differential amplifier, inputs are applied at both the non-inverting and inverting
terminals. The amplifier obtained its name from the fact that the difference between the two input
signals is amplified, hence the difference or differential amplifier.

6
𝑅1
𝑉1
-
𝑉𝑂
+
𝑉2
𝑣𝑖1 𝑅2
𝑅𝐿
𝑣𝑖2

Figure 1.8: Open-loop Differential amplifier

From equation (1.7)


𝑉𝑜 = 𝐴𝑜𝑙 𝑉𝑖𝑛 = 𝐴𝑜𝑙 (𝑉2 − 𝑉1 )
If the two source resistances are neglected,
Then 𝑉1 = 𝑣𝑖1 𝑎𝑛𝑑 𝑉2 = 𝑣𝑖2
Thus 𝑉𝑜 = 𝐴𝑜𝑙 (𝑣𝑖2 − 𝑣𝑖1 ) (1.10)

From equation (1.10), it can be observed that the output, 𝑉𝑜 is 𝐴𝑜𝑙 times the difference between
the two input voltages.

Example 1.1
The op-amp shown in Figure 1.9 is an open-loop differential amplifier with the following
specifications: 𝐴 = 2 × 105 ; 𝑅𝑖𝑛 = 2 𝑀Ω; 𝑎𝑛𝑑 𝑅𝑜𝑢𝑡 = 75 Ω.

𝑅1
-
𝑉𝑂
+
𝑉1 𝑅2
𝑅𝐿
𝑉2

Figure 1.9: An Open-loop operational amplifier for Example 1.1

Determine the output voltage in each of the following cases.


(i) 𝑣𝑖𝑛1 = −7 𝜇𝑉 𝑑𝑐 𝑎𝑛𝑑 𝑣𝑖𝑛2 = 5 𝜇𝑉 𝑑𝑐
(ii) 𝑣𝑖𝑛1 = 20 𝑚𝑉𝑟𝑚𝑠 and 𝑣𝑖𝑛2 = 10 𝑚𝑉𝑟𝑚𝑠

Solution 1.1
(i) 𝐴𝑜𝑙 = 2 × 105 ; 𝑣𝑖𝑛1 = −7 𝜇𝑉 𝑑𝑐; 𝑣𝑖𝑛2 = 5 𝜇𝑉 𝑑𝑐
𝑉𝑜 = 𝐴𝑜𝑙 (𝑉𝑖𝑛2 − 𝑉𝑖𝑛1 )
= 2 × 105 × (5 × 10−6 + 7 × 10−6 )
= 2.4 𝑉

7
(ii) 𝐴𝑜𝑙 = 2 × 105 ; 𝑣𝑖𝑛1 = 20 𝑚𝑉𝑟𝑚𝑠 = 20√2 sin 𝜔𝑡 𝑚𝑉;
𝑣𝑖𝑛2 = 10 𝑚𝑉𝑟𝑚𝑠 = 10√2 sin 𝜔𝑡 𝑚𝑉
𝑉𝑜 = 𝐴𝑜𝑙 (𝑉𝑖𝑛2 − 𝑉𝑖𝑛1 )
= 2 × 105 × (10√2 × 10−3 − 20√2 × 10−3 ) sin 𝜔𝑡
= −2828 sin 𝜔𝑡 𝑚𝑉

1.5 Closed Loop Operational Amplifier Configuration


An open-loop op-amp cannot be used in linear applications. However, an op-amp can effectively
be used in linear applications by introducing feedback from the output to the input. An op-amp
that uses feedback is called a closed-loop amplifier.
The amplifier configuration as observed in Figure 1.10 consists of the op-amp and an external
feedback circuit that connects the output to the inverting input.

+
Vi Vo
-

Negative
feedback
network

Figure 1.10: A Closed-loop operational amplifier

If the signal fed back is out of phase by 180𝑜 with respect to the input, then the feedback is a
negative feedback. On the other hand, if the signal fed back is in phase with that at the input, then
the feedback is a positive feedback.
Negative feedback is one of the most useful concepts in electronics, particularly in operational
amplifier applications. It is a wiring technique where some of the output voltage is sent back to
the inverting terminal. This voltage can be “sent” back through a resistor, capacitor, or complex
circuit or simply can be sent back through a wire.

1.6 Closed-Loop Voltage Gain (Acl)


This is the voltage gain of an op-amp with negative feedback. The closed-loop voltage gain is
determined by the component values in the feedback circuit. This, unlike that of the open-loop
voltage can be controlled by the component values in the feedback circuit.
With negative feedback, the overall closed-loop voltage gain 𝐴𝑐𝑙 , can be reduced and controlled
so that the op-amp can function as a linear amplifier. In addition to providing a controlled, stable
voltage gain, negative feedback also provides for control of the input and output resistances and
amplifier bandwidth.

8
Review Questions
a) What are the connections to a basic op-amp?

b) What are some of the characteristics of a practical op-amp?

c) How does the voltage gain of a practical op-amp differ from an ideal op-amp?

d) What is the op-amp input indicated by +?

e) The op-amp is a high-gain amplifier that has high output resistance and low input
resistance. True/False

f) For an ideal op-amp, the current into each of its two input terminals is zero, and the voltage
across its input terminals is negligibly small. True/False

g) In an amplifier circuit, negative feedback will ………..…… the gain of the amplifier.
i. Increase ii. Decrease iii. Not alter

1.7 Inverting Amplifier


This is an op-amp which is connected in a closed-loop configuration in which the output signal is
applied through a resistor to the inverting input (-).

𝐼𝑓 𝑅𝑓

𝐼𝑖 𝑅𝑖
𝑷 𝑖− 𝑉1
-
𝑉𝑔 𝑉𝑂
𝑖+
𝑉𝑖 +
𝑉2

Figure 1.11: Inverting amplifier

In Figure 1.11, the non-inverting input is grounded. 𝑉𝑖 is connected to the inverting input through
𝑅𝑖 , whilst the feedback resistor, 𝑅𝑓 is connected between the inverting input and output. This
means the output is fed back through 𝑅𝑓 to the inverting input. This represents a negative feedback.
Now, applying the concept of an ideal op-amp, where the input resistance is infinite, this means
there is no current in or out of the inverting input. If there is no current through the input resistance,
then there must be no voltage drop between the inverting and non-inverting inputs.
𝑉𝑖𝑛 = 𝑉2 − 𝑉1 ≃ 0 [refer to equation (1.1)]
Or 𝑉2 = 𝑉1 [refer to equation (1.5)]
It could be said from equation (1.1), that a virtual short-circuit exists between the two terminals.

9
The word ‘virtual’ is used to clarify the fact that, the two input terminals are not actually shorted.
Thus a virtual short circuit means that whatever is the voltage at non-inverting terminal will
automatically appear at the inverting terminal due to the infinite gain.
But the non-inverting input (𝑉2) is grounded and its voltage is zero. Therefore deducing from
equation (1.5), the inverting input terminal (𝑉1) is at ground potential or the voltage is zero (0 V).
This zero voltage at the inverting input terminal is referred to as virtual ground (𝑽𝒈 ). The virtual
ground means that the terminal is not actually connected to the ground, even though, the voltage
at the terminal is zero.

Applying Kirchhoff’s Current Law at point P (Summing Point)


𝐼𝑖 − 𝐼𝑓 − 𝑖− = 0
Since there is no current at the inverting input, then 𝑖− = 0

Thus, the current through 𝑅𝑖 (𝐼𝑖 ) and that through 𝑅𝑓 (𝐼𝑓 ) are equal.
⇒ 𝐼𝑖 = 𝐼𝑓
𝑉𝑖 −𝑉𝑔 𝑉𝑔 −𝑉𝑜
But 𝐼𝑖 = and 𝐼𝑓 =
𝑅𝑖 𝑅𝑓
𝑉𝑖 −𝑉𝑔 𝑉𝑔 −𝑉𝑜
⇒ =
𝑅𝑖 𝑅𝑓

Now 𝑉𝑔 = 0
𝑉𝑖 −𝑉𝑜 𝑉𝑜 −𝑅𝑓
⇒ = or =
𝑅𝑖 𝑅𝑓 𝑉𝑖 𝑅𝑖
𝑉𝑜 𝑅𝑓
∴ 𝐴𝑐𝑙 = =− (1.11)
𝑉𝑖 𝑅𝑖

From equation (1.11),


(i) 𝐴𝑐𝑙 has a negative sign, thus the output has different polarity from that of the input.
(ii) 𝐴𝑐𝑙 depends only on the external resistors

Example 1.2
𝑅𝑓

5 𝑘Ω
- 𝑉𝑂
+
𝑉𝑖

Figure 1.12: for Example 1.2

(a) Given the op-amp configuration in the Figure 1.12, determine the value of 𝑅𝑓 required to
produce a closed-loop voltage gain of -40

10
(b) Calculate:
i. The output voltage, Vo
ii. The current through the 5 kΩ resistor, if Vi = 0.8 V.
(c) If the input remains 5 kΩ, and 𝑅𝑓 is 90 kΩ. What will be the closed-loop voltage gain
produced and the new output voltage?

Solution 1.2
(a) 𝑅𝑖 = 5 𝑘Ω; 𝐴 = −40; 𝑉𝑖 = 0.8𝑉
𝑅𝑓
Now 𝐴 = −
𝑅𝑖
⇒ 𝑅𝑓 = −𝐴𝑅𝑖 = −(−40) × 5
𝑅𝑓 = 200 𝑘Ω

(b)
𝑉𝑜
(i) =𝐴
𝑉𝑖
or 𝑉0 = 𝐴𝑉𝑖
= −40 × 0.8 = - 32 V

𝑉𝑖 −0 0.8
(ii) 𝑖 = =
𝑅𝑖 5000
= 160 µA

(c) 𝑅𝑖 = 5 𝑘Ω; 𝑅𝑓 = 90 𝑘Ω
𝑅𝑓 90
𝐴=− =− = −18
𝑅𝑖 5
𝑉0 = 𝐴𝑉𝑖 = −18 × 0.8
= −14.4 𝑉

1.8 Non-Inverting Amplifier


This is an op-amp that is connected in a closed-loop configuration in which the input signal is
applied to the non-inverting input. A part of the output is applied back to the inverting input
through the feedback circuit. This also represents a negative feedback.

𝑅𝑓
+
𝑉𝑂
- 𝑅𝑖
𝑉𝑖 𝑅𝑓 -
𝑉𝑂
+

𝑅𝑖 𝑉𝑖

Figure 1.13: Non-inverting amplifier


11
From Figure 1.13, the input voltage 𝑉𝑖 is applied directly at the non-inverting input terminal, whilst
the resistor 𝑅𝑖 is connected between the ground and the inverting terminal.
At the inverting terminal,
𝐼𝑖 = 𝐼𝑓
0−𝑉𝑖 𝑉𝑖 −𝑉𝑜
⇒ =
𝑅𝑖 𝑅𝑓
𝑉𝑜 𝑅𝑖 +𝑅𝑓
or =
𝑉𝑖 𝑅𝑖
𝑅𝑓
𝐴 = 1+ (1.12)
𝑅𝑖

From equation (1.12),


(i) A does not have a negative sign, thus the output has the same polarity as the input.
(ii) A depends only on the external resistors

1.9 Voltage Follower (Unity Gain Amplifier)


This is a special case of the non-inverting amplifier where all of the output voltage is fed back to
the inverting input by a straight connection, that is, the output follows the input. In the voltage
follower, the feedback resistor, 𝑅𝑓 = 0 (short circuit), or 𝑅𝑖 = ∞ (open circuit) or both.

-
𝑉𝑂
+

𝑉𝑖

Figure 14: Voltage follower

The straight feedback connection produces a voltage gain of approximately 1, that is


𝑅𝑓
𝐴 = 1+ , but 𝑅𝑓 = 0; or 𝑅𝑖 = ∞
𝑅𝑖
Hence 𝐴 = 1 + 0 = 1
Such a circuit has a very high input resistance and very low output resistance. These features make
it useful as a buffer (intermediate-stage) amplifier to isolate one circuit from another, for example,
a high-resistance from low-resistance load.

Alternatively, as the voltage Vi at the non-inverting by definition is the same as the voltage at the
inverting, then from Figure 1.14, the output voltage, Vo = Vi
𝑉𝑜 𝑉𝑖
But = = =1 (1.13)
𝑉𝑖 𝑉𝑖

12
Example 1.3
60 kΩ

5kΩ
-
Vo

40 mV

Figure 1.15: For Example 1.3


(i) Determine the closed-loop voltage gain of the amplifier
(ii) Find the output of the circuit
(iii) Calculate the current through the feedback resistor.

Solution 1.3

𝑅𝑖 = 5 𝑘Ω; 𝑅𝑓 = 60 𝑘Ω; 𝑉𝑖 = 40 𝑚𝑉
𝑅𝑓 60
(i) 𝐴 = 1+ = 1+ = 13
𝑅𝑖 5

(ii) 𝑉𝑜 = 𝐴𝑉𝑖 = 13 × 40 = 520 𝑚𝑉


𝑉𝑜 −𝑉𝑔 (520−40)×10−3
(iii) 𝑖𝑜 = = = 8𝜇𝐴
𝑅𝑓 60×103

1.10 Comparator
It is a device which is used to sense when a varying signal reaches some threshold value. The
comparator compares two input voltages or signals and produces an output in either of two states
indicating the greater than or less than relationship of the inputs. For example, a comparator can
be used to determine when an input voltage reaches or exceeds a certain defined level; or indicate
whether or not a pulse has an amplitude greater than a particular value. The circuit for the
comparator as seen in Figure 1.16 is the simplest or basic circuit, since no additional external
components are needed.

-
𝑉𝑂
+

𝑉𝑖

Figure 1.16: The comparator

13
In this circuit, the inverting input (-) is grounded to produce a zero-level, whilst the input signal
voltage is applied to the non-inverting input (+).
A very small difference voltage between the two inputs results in the output voltage going to its
limits (very large Vo). This is as a result of the high open-loop voltage gain.
For example, voltage difference, Vdiff (Vin) = 0.20 mV, Aol = 100,000
𝑉𝑜
Since 𝐴𝑜𝑙 = ,
𝑉𝑖

Then 𝑉𝑜 = 100,000 × 0.20 × 10−3 = 20 𝑉

Again, for 741 IC op-amp, if A is differential voltage gain of the op-amp, the minimum input
𝑉𝑠𝑎𝑡
voltage that produces saturation is 𝑉𝐼𝑁 = .
𝐴
for a range of ±15 𝑉, and 𝐴 = 100,000, 𝑉𝑠𝑎𝑡 = ±13.5 𝑉
13.5
∴ 𝑉𝐼𝑁 = = 135 𝜇𝑉
100000

In effect, the comparator can detect very small changes, and thus can be said to be used in
comparing two signals.

t
Vi 0 t

(a) Sinusoidal input voltage

+Vo(max)

Vo t

-Vo(max)

(b) Square output voltage

Figure 1.17: The comparator as a zero-level detector

When a sinusoidal input voltage is applied to the non-inverting input, Figure 1.17 depicts the
outcome. Positive part of the sinewave results in an output of maximum positive level whilst
negative sinewave results in an output of maximum negative level.

14
1.10.1 Non-zero Level Detection
This is a modification to Figure 1.16, and can be used to detect positive and negative voltages.

Vref

Figure 1.18: Fixed reference voltage connected to inverting input

This is done by connecting a fixed reference voltage 𝑉𝑟𝑒𝑓 to the inverting input. When the input
voltage Vin is lower than 𝑉𝑟𝑒𝑓 , the output is at the maximum negative level but goes to the
maximum positive state when Vi exceeds 𝑉𝑟𝑒𝑓 .

𝑉𝑟𝑒𝑓
0
Vin 0
t

+Vo(max)

Vo 0 t
-Vo(max)

Figure 1.19: The Comparator as a Non-zero Level Detector

1.11 Summing Amplifier (Adder or Summer)


It is a circuit that combines several inputs and produces an output voltage which is proportional
to the negative of the algebraic sum of its input voltages.
𝑅𝑓
𝑅1
𝑉1
𝑅2
𝑉2
𝑂
-
𝑉3 𝑉𝑂
𝑅3
+
𝑉4
𝑅4
Figure 1.20: Summing Amplifier with Four Inputs
15
The summing amplifier is a variation of the inverting amplifier, and it is taking the advantage of
the fact that the inverting configuration can handle many inputs at the same time.

Since point O is a virtual ground just as in an inverting amplifier, then current entering each op-
amp is zero.

Thus
𝑉1 −𝑉𝑔 𝑉2 −𝑉𝑔 𝑉3 −𝑉𝑔 𝑉4 −𝑉𝑔 𝑉𝑔 −𝑉𝑜
𝑖1 = , 𝑖2 = , 𝑖3 = , 𝑖4 = 𝑖𝑓 =
𝑅1 𝑅2 𝑅3 𝑅4 𝑅𝑓

Applying KCL, 𝑖𝑓 = 𝑖1 + 𝑖2 + 𝑖3 + 𝑖4
𝑉𝑔 −𝑉𝑜 𝑉1 −𝑉𝑔 𝑉2 −𝑉𝑔 𝑉3 −𝑉𝑔 𝑉4 −𝑉𝑔
= + + +
𝑅𝑓 𝑅1 𝑅2 𝑅3 𝑅4
But 𝑉𝑔 = 0
−𝑉𝑜 𝑉1 𝑉2 𝑉3 𝑉4 𝑉𝑜 𝑉1 𝑉2 𝑉3 𝑉4
⇒ = + + + or = −( + + + )
𝑅𝑓 𝑅1 𝑅2 𝑅3 𝑅4 𝑅𝑓 𝑅1 𝑅2 𝑅3 𝑅4
𝑅 𝑅 𝑅 𝑅
𝑉𝑜 = − (𝑅𝑓 𝑉1 + 𝑅𝑓 𝑉2 + 𝑅𝑓 𝑉3 + 𝑅𝑓 𝑉4 ) (1.14)
1 2 3 4

NB:
If the four input resistors are the same, that is 𝑅1 = 𝑅2 = 𝑅3 = 𝑅4 = 𝑅
𝑅𝑓
𝑉𝑜 = − (𝑉1 + 𝑉2 + 𝑉3 + 𝑉4 ) (1.15)
𝑅

= −𝐴(𝑉1 + 𝑉2 + 𝑉3 + 𝑉4 )
Thus the output voltage is proportional to the negative of the sum of the input voltages.
But if 𝑅𝑓 = 𝑅1 = 𝑅2 = 𝑅3 = 𝑅4 = 𝑅, then the output voltage
𝑅
𝑉𝑜 = − 𝑅 (𝑉1 + 𝑉2 + 𝑉3 + 𝑉4 ))
𝑉𝑜 = −(𝑉1 + 𝑉2 + 𝑉3 + 𝑉4 )) (1.16)
Hence the output is exactly equal to the sum of the input voltages.

Example 1.4
Calculate 𝑉𝑂 and 𝑖𝑓 in the op-amp circuit in Figure 1.21.

5 𝑘Ω 20 𝑘Ω

- 𝑉𝑂
3.5 𝑉
4 𝑘Ω +

6𝑉

Figure 1.21 for Example 1.4


16
Solution 1.4
For a summer of two inputs
20 20
𝑉𝑜 = − [ 5 (3.5) + (6)] = −44 𝑉
4

0−𝑉𝑜 44
𝑖𝑓 = = = 2.20 𝑚𝐴
𝑅𝑓 20×103
Note: Specifically for the example above, 𝑖0 = −𝑖𝑓

Example 1.5
Find 𝑉𝑜 𝑎𝑛𝑑 𝑖𝑜 in the op-amp circuit shown in Figure 1.22.
15 𝑘Ω 30 𝑘Ω

12 𝑘Ω
- 𝑖𝑜
𝑉𝑂
6 𝑘Ω +
1.5 𝑉
−2 𝑉 10 𝑘Ω
1.2 𝑉

Figure 1.22 is for Example 1.5

Solution 1.5
For a summer of three inputs
30 30 30
𝑉𝑜 = − [ (1.5) + (−2) + (1.2)] = −4 𝑉
15 12 6
Current 𝑖𝑜 is the sum of the currents through the 30 kΩ and 10 kΩ resistors. Both resistors have
voltage, 𝑉𝑜 = −6 𝑉 across them, since 𝑉𝑔 = 0

𝑉𝑜 −0 𝑉𝑜 −0
Hence 𝑖𝑜 = + = −0.133 − 0.4 = −0.533 𝑚𝐴
30×103 10×103

1.12 Cascaded Op-Amp Circuits


Since op-amp circuits are modules or building blocks for designing complex circuits, it is often
necessary in practical applications to connect the circuits in cascade. This is a head-to-tail
arrangement of two or more op-amp circuits such that the output of one is the input of the next.

+ +
V1
Stage 1 Stage 2 V3 = A2V2 Stage 3
V2 = A1V1 Vo = A3V3
- A1 A2 A3 -

Figure 1.23: A three – stage cascaded connection


17
When the circuits are cascaded, each circuit in the string is called a stage. The original input signal
is increased by the gain of the individual stage.

At the end of stage 3,


The output Vo, with respect to the input, V1 can be derived from:
𝑉𝑜 = 𝐴3 𝑉3 = 𝐴3 (𝐴2 𝑉2 ) = 𝐴3 𝐴2 (𝑉2 ) = 𝐴3 𝐴2 (𝐴1 𝑉1 ) = (𝐴3 𝐴2 𝐴1 )𝑉1 (1.17)

Example 1.6
Find Vo and io in the circuit in Figure 1.24.
12 𝑘Ω 𝑖𝑂
15 𝑘Ω
7 𝑘Ω
6 𝑘Ω -
𝑏 𝑉𝑂
𝑎
-
+
+

4 mV

Figure 1.24: for Example 1.6

Solution 1.6
The circuit consists of two non-inverting op-amps cascaded.
15
At the output of the first op-amp: 𝑉𝑜𝑎 = (1 + ) × 4 = 14 𝑚𝑉
6
12
At the output of the second op-amp: 𝑉𝑜 = (1 + 7
) × 14 = 38 𝑚𝑉

Example 1.7
For the circuit in figure 1.25, find 𝑉𝑜 .
25 𝑘Ω
40 𝑘Ω
100 𝑘Ω
20 𝑘Ω
20 𝑘Ω
- -
+
6𝑉 10 𝑘Ω +

4𝑉
𝑉𝑂
2𝑉

Figure 1.25: Circuit for Example 1.7

18
Solution 1.7
Starting with the 4 − 𝑉 input, its output, 𝑉04 is given by
𝑅𝑓4 40
𝑉04 = − ×4𝑉 =− × 4 = −8 𝑉
𝑅4 20
This then leads to a 3-input summing amplifier.
Hence the output voltage 𝑉0 can be obtained using
100 100 100
𝑉0 = − [ 25 × 6 + × (−8) + × 2]
20 10

= −[24 − 40 + 20] = −4 𝑉

1.13 Difference Amplifier (Subtractor)


This is a circuit whose output is proportional to the difference of two input signals. In other words,
the difference amplifier amplifies the difference between two inputs but rejects any signals
common to the two inputs. It is also sometimes known as the subtractor.
𝑅𝑓

𝑅𝑖
-
𝑉𝑂
+
𝑉1
𝑉2

Figure 1.26: Difference amplifier

Considering the op-amp circuit in figure 1.26, inputs are applied to both the inverting and the non-
inverting terminals.
By applying the principle of superposition, the output
𝑉𝑜 = 𝑉𝑜1 + 𝑉𝑜2 (1.18)
Where 𝑉𝑜1 is the output produced by 𝑉 1 and 𝑉𝑜2 is the output produced by 𝑉2
Thus for the inverting terminal,
𝑅
𝑉𝑜1 = − 𝑅𝑓 𝑉1
𝑖

For the non-inverting


𝑅𝑓
𝑉𝑜2 = (1 + ) 𝑉2
𝑅𝑖

Since 𝑉𝑜 = 𝑉𝑜1 + 𝑉𝑜2


𝑅 𝑅𝑓
= − 𝑅𝑓 𝑉1 + (1 + ) 𝑉2
𝑖 𝑅𝑖
𝑅𝑓 𝑅
or 𝑉𝑜 = (1 + ) 𝑉2 − 𝑅𝑓 𝑉1 (1.19)
𝑅𝑖 𝑖

19
𝑅𝑓
Now if 𝑅𝑓 ≫ 𝑅𝑖 , then ≫1
𝑅𝑖
𝑅𝑓 𝑅𝑓
Thus 𝑉𝑜 = 𝑉2 − 𝑉1
𝑅𝑖 𝑅𝑖
𝑅𝑓
or 𝑉𝑜 = (𝑉2 − 𝑉1 ) (1.20)
𝑅𝑖

Now, if 𝑅𝑓 = 𝑅𝑖 (from 1.20)


Then 𝑉𝑜 = 𝑉2 − 𝑉1
Thus the difference amplifier becomes a subtractor.

Figure 1.27 is a variation of Figure 1.26 with introduction of additional resistors to the non-
inverting side of the op-amp.
𝑅2

𝑅1
𝑎
-
𝑏 𝑉𝑂
+
𝑉1 𝑅3 = 𝑅1
𝑉2 𝑅4 = 𝑅2

Figure 1.27: Difference amplifier

Let 𝑉𝑎 and Vb be the voltages at nodes a and b respectively.


Applying KCL to node a,
𝑉1 −𝑉𝑎 𝑉𝑎 −𝑉𝑜
=
𝑅1 𝑅2
𝑅2 𝑅2
⟹ 𝑉𝑜 = ( + 1) 𝑉𝑎 − 𝑉
𝑅1 𝑅1 1

Applying KCL to node b,


𝑉2 −𝑉𝑏 𝑉𝑏 −0 𝑅4
= or 𝑉𝑏 = 𝑉
𝑅3 𝑅4 𝑅3 +𝑅4 2

But 𝑉𝑎 = 𝑉𝑏
𝑅 𝑅 𝑅2
⟹ 𝑉𝑜 = (𝑅2 + 1) (𝑅 +𝑅
4
) 𝑉2 − 𝑉 (1.21)
1 3 4 𝑅1 1

Now if 𝑅1 = 𝑅3 and R 2 = R 4
𝑅2
Then 𝑉𝑜 = (𝑉2 − 𝑉1 )
𝑅1

And if 𝑅1 = 𝑅2
Then 𝑉𝑜 = 𝑉2 − 𝑉1

20
Example 1.8
If in a difference amplifier circuit, 𝑅1 = 10 𝑘Ω; 𝑅2 = 20 𝑘Ω; 𝑉1 = 5 𝑉 and 𝑉2 = 6 𝑉, find the
value of the output voltage.

Solution 1.8
𝑅2 𝑅2
𝑉𝑜 = (1 + ) 𝑉2 − 𝑉
𝑅1 𝑅1 1
20 20
= (1 + 10) 6 − (10 × 5) = 8 𝑉

Example 1.9
The circuit in Figure 1.28 is a difference amplifier. Find 𝑣𝑜 given that 𝑣1 = 5 𝑉 and 𝑣2 = 10 𝑉.
20 𝑘Ω

5 𝑘Ω
-
+
𝑉1 10 𝑘Ω
10 𝑘Ω 𝑉𝑂
𝑉2

Figure 1.28: for Example 1.9

Solution 1.9
For the difference amplifier,
𝑅 𝑅 𝑅2
𝑉𝑜 = (𝑅2 + 1) (𝑅 +𝑅
4
) 𝑉2 − 𝑉
1 3 4 𝑅1 1

Where 𝑅1 = 5 𝑘Ω; 𝑅2 = 20 𝑘Ω; 𝑅3 = 10 𝑘Ω; 𝑅4 = 10 𝑘Ω; 𝑣1 = 5 𝑉 𝑎𝑛𝑑 𝑣2 = 10 𝑉.


20 10 20
Thus 𝑉𝑜 = ( 5 + 1) (10+10) × 10 − ×5
5
=5𝑉

1.14 Op-Amp Integrator


The op-amp integrator is a circuit whose function is to produce an inverted output (voltage) which
is proportional to the area under the curve of the input function (voltage). This is the integral of
the input voltage.
From Figure 1.29, it could be observed that the integrator is similar to that of the inverting
amplifier except that the feedback component is a capacitor C instead of a resistor 𝑅𝑓 . The
capacitor forms an RC circuit with the input resistor.

21
𝐶
𝐼𝐶
𝑅
-
𝐼𝑅 𝑉𝑂
+
𝑉𝑖

Figure 1.29: An ideal op-amp integrator

Just like the inverting amplifier, the inverting input is at virtual ground, thus input current through
𝑅𝑖 (𝐼𝑖 ) is the same as Ic through C.
Now
𝑉𝑖 −0 𝑉𝑖
𝐼𝑅 = =
𝑅𝑖 𝑅𝑖

The displacement current relation is used to find Ic


That is
𝑑𝑉 𝑑(𝑉𝑔 −𝑉𝑜 )
𝐼𝑐 = 𝐶 =𝐶
𝑑𝑡 𝑑𝑡
𝑑(0−𝑉𝑜 ) 𝑑𝑉𝑜
𝐼𝑐 = 𝐶 = −𝐶
𝑑𝑡 𝑑𝑡
But 𝐼𝑅 = 𝐼𝐶
𝑉𝑖 𝑑𝑉𝑜 1
⟹ = −𝐶 or 𝑑𝑉𝑜 = − 𝑅 𝐶 𝑉𝑖 𝑑𝑡
𝑅𝑖 𝑑𝑡 𝑖

1
Thus 𝑉𝑜 = − 𝑅 𝐶 ∫ 𝑉𝑖 𝑑𝑡 (1.22)
𝑖

As could be observed in equation (1.22), the input signal is integrated at the output.

Example 1.10
The integrator in Figure 1.29 has resistance, 𝑅 = 50 𝑘Ω and capacitance, 𝐶 = 5 𝜇𝐹. Determine
the output voltage when a dc voltage of 20 mV is applied at time, 𝑡 = 0.

Solution 1.10
𝑅 = 50 𝑘Ω; 𝐶 = 5 𝜇𝐹; 𝑉𝑖 = 20 𝑚𝑉 𝑎𝑡 𝑡 = 0
1
𝑉𝑜 = − 𝑅 𝐶 ∫ 𝑉𝑖 𝑑𝑡
𝑖
1 𝑡
= − ∫0 20𝑑𝑡 = −80𝑡 𝑚𝑉
50×103 ×5×10−6

Example 1.11
If 𝑉1 = 20 cos 4𝑡 𝑚𝑉, 𝑉2 = 10 sin 2𝑡 𝑚𝑉 and 𝑉3 = 5𝑡 mV, find Vo in the op-amp circuit in
figure 1.30. Assume that the voltage across the capacitor is initially zero

22
2 𝑀Ω 10 𝜇𝐹

200 𝑘Ω
-
250 𝑘Ω +
𝑉1
𝑉2

𝑉3

Figure 1.30: for Example 1.11


Solution 1.11
This is a summing integrator and
1 1 1
𝑉𝑜 = − ∫ 𝑉1 𝑑𝑡 − ∫ 𝑉2 𝑑𝑡 − ∫ 𝑉3 𝑑𝑡
𝑅1 𝐶 𝑅2 𝐶 𝑅3 𝐶

1 𝑡 1 𝑡
=− −6 ∫0 20 cos 4𝑡 𝑑𝑡 −
6 5 −6 ∫0 10 sin 2𝑡 𝑑𝑡 −
2×10 ×10×10 2×10 ×10×10
1 𝑡
∫ 5𝑡 𝑑𝑡
250×103 ×10×10−6 0
1 5
= − 4 sin 4𝑡 + 2 cos 2𝑡 − 𝑡 2 𝑚𝑉

1.15 Op-Amp Differentiator


The op-amp differentiator is a circuit that produces an inverted output signal (voltage) which is
proportional to the rate of change of the input signal (voltage).
𝑅
𝐼𝑅
𝐼𝐶 𝐶
-
𝑉𝑂
+
𝑉𝑖

Figure 1.30: An ideal op-amp differentiator


The differentiator is an inverse mathematical operation to that of the integrator.
In the circuit, the capacitor and the resistor are interchanged, thus making the capacitor the input
element.

From Figure 1.30,


Because of virtual ground, 𝑉𝑔 = 0

23
Also, since the current to the input of the op-amp is zero, then
𝐼𝑐 = 𝐼𝑅
Where Ic = current through capacitor and IR = current through feedback resistor.
But Ic, using the displacement current equation is
𝑑𝑉
𝐼𝑐 = 𝐶 , where 𝑉 = 𝑉𝑖 − 𝑉𝑔
𝑑𝑡
𝑑 𝑑
= 𝐶 𝑑𝑡 (𝑉𝑖 − 0) = 𝐶 𝑑𝑡 𝑉𝑖
𝑉𝑔 −𝑉𝑜 −𝑉𝑜
and 𝐼𝑅 = =
𝑅 𝑅
−𝑉𝑜 𝑑
Thus =𝐶 𝑉
𝑅 𝑑𝑡 𝑖
𝑑
or 𝑉𝑜 = −𝑅𝐶 𝑉 (1.23)
𝑑𝑡 𝑖

Thus for the differentiator as could be seen from equation (1.23), the input signal is differentiated
at the output.

Example 1.11
Find Vo in Figure 1.31 if Vi is a sinusoidal voltage of peak value 10 mV, and 𝑓 = 500 𝐻𝑧.

100 kΩ

5 𝜇𝐹
-
𝐶 𝑉𝑂
+
𝑉𝑖

Figure 1.31: for example 1.11

Solution 1.11
𝑉𝑖 = 𝐴 sin 𝜔 𝑡 = 𝐴 sin 2𝜋𝑓 𝑡
= 10 sin 2𝜋 × 500 𝑡 = 10 sin 1000𝜋 𝑡
𝑑
But 𝑉𝑜 = −𝑅𝐶 𝑑𝑡 𝑉𝑖
𝑑
= −105 × 5 × 10−6 𝑑𝑡 (10 sin 1000𝜋𝑡)
𝑉𝑜 = −5000𝜋 cos 1000𝜋 𝑡 𝑚𝑉

24
1.16 Problems
P.1. Calculate the voltage gain of Figure 1.32 [11]

Figure 1.32: for P. 1

P.2. Calculate the input voltage, V1, if 𝑅1 = 100 Ω; 𝑅𝑓 = 1 𝑘Ω, and 𝑉𝑜𝑢𝑡 = 550 𝑚𝑉

Figure 1.33 for P2 [50 mV]

P.3 Calculate the output voltage in Figure 1.34, if R1 = R2 = R3 = 100 , Rf = 1 k , and V1 =


V2 = V3 = 50 mV.

Figure 1.34 for P3 [-1.5 V]

P.4 Calculate the output voltage in Figure 1.35 if 𝑉1 = 𝑉2 = 700 𝑚𝑉.

Figure 1.35 for P4 [0 V]

25
P.5 Calculate the input voltage if the final output, Vo, is 20.40 V.

Figure 1.36 for P5 [0.34 V]

P. 6 For the circuit in Figure 1.37, determine the value of 𝑉2 in order to make 𝑉𝑜 = −16.5𝑉.

10 𝑘Ω 50 𝑘Ω
2𝑉
20 𝑘Ω
𝑉2 -
𝑉𝑂
50 𝑘Ω +
−1 𝑉

Figure 1.37 for P6

P. 7 For the circuit in Figure 1.38, determine the output voltage given that 𝑉1 = 1 𝑉 and 𝑉2 =
1 𝑉.
30 𝑘Ω

2 𝑘Ω
-
𝑉𝑂
+
𝑉1 2 𝑘Ω
𝑉2 20 𝑘Ω

Figure 1.38 for P7

26
Assignment
Q1. Find 𝑉𝑜 𝑎𝑛𝑑 𝑖𝑜 in the differential amplifier in Figure A1.
2 𝑘Ω 4 𝑘Ω

- 𝑖𝑂
1 𝑘Ω
+
10 𝑉 +
3 𝑘Ω 5 𝑘Ω 𝑉𝑂
8𝑉

Figure A1 for Q1

Q2.
(i) An op-amp integrator has 𝑅 = 100 𝑘Ω 𝑎𝑛𝑑 𝐶 = 100 𝑛𝐹. If the input voltage is
𝑣𝑖 = 20 cos 100𝑡 𝑚𝑉, obtain the output voltage.
(ii) If the positions of R and C are interchanged, determine the output.

Q3.
Determine 𝑉𝑜

2.4 𝑘Ω 48 𝑘Ω
45 𝑘Ω

3.5 𝑉 6 𝑘Ω
- 25 𝑘Ω
-
16 𝑘Ω + 𝑉𝑂
2.5 𝑉 +
36 𝑘Ω
−1.5 𝑉 24 𝑘Ω

24 𝑘Ω 60 𝑘Ω

−3 𝑉
-
25 𝑘Ω +

4𝑉

Figure A2 for Q3

27
CHAPTER 2
FEEDBACK AND STABILITY
2.0 Introduction
Control systems contribute to every aspect of modern society. They have widespread applications in
science and industry, from steering ships and planes to guiding missiles and the space shuttle. They
could also be found in domestic applications such as: electric iron thermostat, refrigeration control,
WC tank, water level control and hot water heater control.
Control systems also exist naturally; our bodies contain numerous control systems.

Control systems analysis and design focuses on three main primary objectives:
i. Producing the desired transient response
ii. Reducing steady-state errors
iii. Achieving stability
Transient response is the sum of the natural response and forced responses, but the natural response
is large. In steady-state, the response is also the sum of the natural response and forced responses,
but the natural response is small. Steady-state response determines the accuracy of the control system;
it governs how closely the output matches the desired response.
A system must be stable in order to produce the proper transient and steady-state response.
Control systems can be open-loop or closed-loop.

2.1 Open-Loop System


The open-loop control system is also known as control system without feedback or non-feedback --
control system. It is a system that cannot compensate for any disturbances that add to the controller’s
driving signal. It does not monitor or correct the output for disturbances. In this system, the control
action is independent of the desired output. Thus, the output is not compared with the reference input.
Disturbance 1 Disturbance 2

+ + Output or
Input or Input Process or
Controller Controlled
Reference Transducer + Plant + Variable
Summing Summing
Junction Junction
Figure 2.1: An Open-loop System

A generic open-loop system is shown in Figure 2.1. It starts with a subsystem called the input
transducer, which converts the input to that used by the controller. The controller drives a process
or plant. The input is sometimes called the reference, while the output can be called the controlled
variable. Other signals, such as disturbances, are added to the controller and process outputs via
summing junctions, which yield the algebraic sum of their input signals using associated signs.

The output of an open-loop system is corrupted not only by signals that add to the controller’s
commands (disturbance 1) but also disturbances at the output (disturbance 2) of which the system
cannot correct. Open-loop systems are commanded by the input.
Mechanical systems consisting of a mass, spring and damper with a constant force positioning the
mass are open-loop systems. The system position will change with disturbance. The disadvantages
of this system are: sensitivity to disturbances and inability to correct these disturbances.
28
Advantages
a. They are simple.
b. They are economical.
c. Less maintenance is required and not difficult.
d. Proper calibration is not a problem.

Disadvantages
a. Open-loop systems are inaccurate
b. They are not reliable.
c. They are slow.

2.2 Closed-Loop (Feedback Control) System


Technically, feedback means transferring a portion of the energy from the output of some device back
to its input. It may be defined as a process of injecting some energy from the output and then return
it back to the input. In other words, when a part or fraction of output is combined to the input,
feedback is said to exist.
It makes the input dependent on the output, and makes it possible to control any process. In other
words the control action is dependent on the desired output. Feedback makes it possible to “spy” on
the output of some process and thereby control the input to the process in accordance with the
effectiveness of the output. This kind of arrangement is termed as a closed-loop control system;
because the input depends on feedback from the output.
Thus closed-loop control systems are also known as feedback control systems. Closed-loop systems
monitor the output and compare it to the input. If an error is detected, the system corrects the output
and hence the effects of disturbance. Generally, it is the most effective control system possible.

A generic architecture of a closed-loop system is shown in Figure 2.2.

Error or Disturbance 1 Disturbance 2


Actuating
Input or Input + Signal + +
+ Output or
Controller + Process Controlled
Reference Transducer or Plant
- Variable

Output
Transducer
or Sensor

Figure 2.2: A Closed-loop


System
Here, an output transducer or sensor measures the output response and converts it into the form used
by the controller.
Feedback path is the return path from the output to the summing junction. In Figure 2.2, the output
signal is subtracted from the input signal and the result is generally called the actuating signal.
However, in systems where both the input and output transducers have unity gain (i.e. the transducer
amplifies its input by 1), the actuating signal’s value is equal to the actual difference between the
input and the output. Under this condition, the actuating signal is called the error.

The closed-loop system compensates for disturbances by measuring the output response, feeding that
measurement back through a feedback path, and comparing that response to the input at the summing
29
junction. The system drives the plant through the actuating signal to make a correction. If there is no
difference, the system does not drive the plant since the plant response is already the desired response.

In summary, systems that perform the previously described measurement and correction are called
closed-loop or feedback control systems. Systems that do not have this property of measurement and
correction are called open-loop systems.

Advantages of Closed Systems


a. These systems are more reliable.
b. Closed loop systems are faster.
c. A number of variables can be handled simultaneously.

Disadvantages
a. Closed loop systems are expensive
b. Maintenance difficult
c. Complicated installation

2.3 Feedback Relationships and Definitions


What are amplifiers?
The purpose of an amplifier is to increase the power in a signal without distorting it. This process is
called amplification. For example, amplifiers are used to increase the power in the signal from a
microphone so that it can drive a loudspeaker, or to amplify the control signals produced by the pilot
of an aircraft so that they will move the flaps on the wings and other control devices. The extra power
must come from a power supply. In the case of an electronic amplifier, the power supply gets its
power from the ac supply mains or a battery.

The input signal causes the amplifier to control the flow of current from the voltage supply to the
load. Thus, more power may be delivered to the load than is taken from the input signal source. In
practice, amplification usually means increasing the voltage amplitude of the signal into the given
load.
An ideal linear amplifier in the mid-frequency region provides an output signal that is an exact replica
of the applied input signal but the practical amplifiers depart from the ideal one for several reasons
like non-linearity of transistor characteristics, parameter variation, temperature effects etc. Some of
these factors can be minimised by improving the involved basic devices. However, there is a limit,
beyond which this approach does not work. The best approach is to use the principle of feedback to
achieve a desired degree of improvement.
The gain of an amplifier without feedback is subject to modification due to a range of external factors.
These include change of dc supply levels, change of component values of device parameters (due to
ageing or replacement), variation in components and devices in a production spread, change in
external loading and change in environmental conditions. The corresponding changes in gain for a
feedback amplifier will be greatly reduced. A typical reduction could be from a 20% variation in the
basic system gain to a 1% variation in the feedback system.
Further properties of feedback, which are of particular interest in amplifiers, include the reduction of
harmonic distortion due to non-linearity within the amplifier and modification of both input and
output impedances. Again, it makes amplifier operation more stable in respect to variations in gain
due to line voltage changes, tube differences, aging, etc.

30
2.4 Types of Feedback
Depending upon whether the feedback signal increases or decreases the input signal, there are two
basic types:
a) Regenerative or direct or positive feedback and
b) Degenerative or inverse or negative feedback

2.4.1 Regenerative (Positive) feedback


This takes place when the feedback signal (voltage or current) is in phase with the applied signal, and
thus aids it. In this case the feedback is so applied as to increase the input signal. Positive feedback
increases the gain and also causes excessive distortion and instability of an amplifier. If the gain is
sufficiently large it leads to oscillations, hence, it is used in oscillator circuits

2.4.2 Degenerative (Negative) feedback


This is the term used when the feedback energy is out of phase with the applied signal and thus
opposes it. This tends to weaken the input signal. Negative feedback decreases the amplifier gain but
it has numerous advantages such as gain stability, reduction in nonlinear distortion, reduction in
noise, increase in bandwidth or improvement in frequency response, increase in input impedance and
reduction in output impedance. Thus subsequently, much time will be spent looking at negative
feedback since it is frequently used in amplifier circuits.

2.4.3 Comparison of Positive and Negative feedback

Table 2.1: Feedback comparison


Parameter Positive Negative
1. 1. Overall phase shift 0 or 360o 180o
2. 2. Feedback signal and input Are in phase Are out of phase
signal
3. 3. Input voltage Increases due to feedback Decreases due to feedback
4. 4. Output voltage Increases due to feedback Decreases due to feedback
5. 5. Voltage gain Increases due to feedback Decreases due to feedback
6. Stability Gets poorer as feedback increases Gets better as feedback increases
7. Application Oscillators Amplifiers
8. Noise Increases with feedback Decreases with feedback

2.4.4 Feedback Relations


The properties of feedback amplifiers and closed-loop control systems are so closely linked that it is
useful to compare the basic definitions and the resulting expression for the system gain. Figure 2.3
and 2.4 (page 34) show block diagrams for the two system types.

A feedback amplifier may be defined as an amplifier for which the terminal input signal is the sum
of an external signal and a signal proportional to the output signal.
The signal proportional to the output signal is the feedback signal i.e. f  So  f =So
where  is the proportionality constant called the feedback network gain.

31
Terminal
Input Output
External + Amplifier gain
signal A
Si S So
+

Feedback
network gain

Figure 2.3: Feedback amplifier system

Figure 2.3 shows that the signal proportional to the output signal is So and this is added to the
external signal Si at the summing junction. The resulting terminal input signal S is then amplified by
the gain A to give the output So. Signals Si, So and S are used at this stage as in practical systems, and
may be considered as either voltage signals or current signals according to the method of connection.

From the definition of the feedback amplifier which is shown in figure 2.3;
S  = Si + So (2.1)
So = AS  (2.2)
So = A(Si + So ) = ASi + ASo
 So − ASo = So (1 − A) = ASi
So A
 Af = =
Si 1 − A

Therefore, the system gain or gain with feedback is:


A
Af = (2.3)
1 − A

Note:
Equation (2.3), was obtained with the assumption that the portion fed back, 𝛽, is positive, and hence
positive feedback or regeneration takes place. The quantity, 𝛽𝐴, represents the amplitude of the
feedback signal (sometimes called feedback factor), superimposed upon 𝑆𝑖 .
The larger this feedback factor, the smaller is the denominator of the above expression and hence the
larger is the gain with feedback. Unfortunately, increasing the gain by positive feedback also
increases the distortion and noise in the same proportion. Thus positive feedback is rarely used except
for oscillators.

For negative feedback, the fraction, 𝛽, and hence 𝛽𝐴 becomes negative (i.e. – 𝛽𝐴), and the gain with
feedback,
𝐴
𝐴𝑓 = (2.4)
1+𝛽𝐴
The greater the feedback factor, the smaller is the feedback gain of the amplifier, but also the smaller
is the distortion introduced by the amplifier.

32
When the feedback factor is made large compared to 1, as is the case for large amounts of feedback,
the denominator of the above expression (2.4), becomes
𝐴 1
𝐴𝑓 = = (2.5)
𝛽𝐴 𝛽
The gain of the amplifier in this case is quite small, but depends only on the feedback fraction, β, and
is thus substantially independent of the actual gain, A, of the amplifier.

Example 2.1
1
An amplifier with voltage gain of 80 𝑑𝐵 uses 25 of its output in negative feedback. Determine the
gain with feedback in dB.

Solution 2.1
1
Open-loop voltage gain, 𝐴 = 80 𝑑𝐵 𝑜𝑟 1080⁄20 = 10000; feedback ratio, 𝛽 = 25 = 0.04
𝐴
But gain with feedback (negative), 𝐴𝑓 =
1+𝛽𝐴
10000
=
1+0.04×10000
= 24.94
𝐴𝑓 in dB = 27.94 dB (That is, 20 log 24.94)

Example 2.2
A single stage transistor amplifier has a voltage gain of 500 without feedback and 50 with feedback.
Find the percentage of output which is fedback to the input side.

Solution 2.2
𝐴 = 500; 𝐴𝑓 = 50
𝐴
Now 𝐴𝑓 =
1+𝛽𝐴
500
50 = 1+500𝛽

Thus 𝛽 = 0.018 or 1.8%

2.4.5 Amplifier Gain A, and  as Complex Quantities


Both gain A and feedback factor  may be complex quantities having both modulus and angle each
of which depends upon signal frequency. A more general form of equation (2.3) includes these as
A and . The resulting expression is then:

A
Af = (2.6)
1 − A( +  )

For a particular amplifier, the modulus of the denominator 1 − A( +  ) may either be greater
than one or less than one depending on the signal frequency. These two conditions in turn result in a
decrease or an increase respectively in the gain compared with the gain without feedback. For
convenience, the two conditions are known as negative and positive feedback and may be defined in
the following way:

33
The feedback is negative if the modulus of the gain with feedback |𝐴𝑓 | is less than the gain without
feedback |𝐴|.
The feedback is said to be positive if the modulus of the gain with feedback |𝐴𝑓 | is greater than the
gain without feedback.|𝐴|
i.e. 𝐴𝑓    A negative feedback
𝐴𝑓    A positive feedback.

Many of the properties due to feedback are most marked in the frequency ranges for which
(+)=.
The resulting gain equation becomes
A
Af = (2.7)
1 + A
and the condition can be referred to as simple negative feedback, SNFB. The different types of
feedback can be illustrated by the example 2.3

Example 2.3
The gain and feedback factor of an amplifier have the different values at different frequencies listed
in the table 2.1 below. In each case, determine the nature of the feedback and the system gain.

Table 2.2: Values of A and 𝛽 for example 2.3


𝑓1 𝑓2 𝑓3 𝑓4
A 5000180 4500160 100065 50020
 0.020 0.018-5 0.001148-10 0.001-15

Solution 2.3
At frequency, 𝑓1 , the gain A is 5000180 and the feedback factor  is 0.020. Applying equation
(2.8), we have:

𝐴∠𝜃 5000∠180𝑜
𝐴𝑓 = =
1−𝛽𝐴∠(𝜃+∅) 1−[5000×0.02∠(180𝑜 +0𝑜 )]
5000∠180𝑜 5000∠180𝑜
= 𝑜 =
1−(100 cos 180𝑜 +𝑗100 sin 180 1−(−100)
5000∠180𝑜
𝐴𝑓 = = 49.5∠180o
101

This is an example of simple negative feedback since  +  = 180

For f2,
4500160 4500160 4500160
Af = = =
1 − 81155 1 − (− 73.4 + j34.2) 81.9 − 24.7
54.9∠185o
The feedback is again negative since the gain with feedback, 54.9 is less than the gain without
feedback 4500.

34
For f3,
100065 100065 100065
Af = = =
1 − 1.14855 1 − (0.658 + j 0.94) 1 − 70
= 1000∠135o
In this case, the gain modulus is unchanged so the feedback is neither positive nor negative. It has
however, had the effect of changing the phase angle.

Finally at f4,
50020 50020 50020
Af = = =
1 − 0.55 1 − (0.498 + j 0.044) 0.503 − 5
= 992∠25o

Thus at f4, the feedback is positive since 992 is greater than the gain of 500 without feedback at this
frequency.

2.5 Closed-Loop System Gain


As already explained, a closed loop system is one in which output is fed back into an error detector
and compared with the reference input. In the equivalent control system for figure 2.2, the error signal
is the difference between a reference input and the processed output. The use of the difference instead
of the sum used for the feedback amplifier causes a sign change between the formula for system gain
for the amplifier and that for the control system.

Consider a closed loop system shown in Figure 2.4


where, R(s) = Reference input
E(s) = Actuating signal or error signal
G(s) = Forward path transfer function
C(s) = Output signal
H(s) = Feedback transfer function
B(s) = Feedback signal

Error
Reference + signal System gain Output
Input, R(s) E(s)
G(s)
C(s)
--
B(s)
Sensing system
gain
H(s)

Figure 2.4: Closed-loop control system

From Figure 2.4


𝐶(𝑠) = 𝐺(𝑠)𝐸(𝑠) (2.10)
𝐵(𝑠) = 𝐻(𝑠)𝐶(𝑠) (2.11)

35
𝐸(𝑠) = 𝑅(𝑠) − 𝐵(𝑠) (2.12)

Putting E(s) in equation (2.12) into equation (2.10)


𝐶(𝑠) = 𝐺(𝑠)[𝑅(𝑠) − 𝐵(𝑠)]
= 𝐺(𝑠)𝑅(𝑠) − 𝐺(𝑠)𝐵(𝑠)
= 𝐺(𝑠)𝑅(𝑠) − 𝐺(𝑠)𝐻(𝑠)𝐶(𝑠)

Therefore the closed loop transfer function is given as:


𝐶(𝑠) 𝐺(𝑠)
= (2.13)
𝑅(𝑠) 1+𝐺(𝑠)𝐻(𝑠)
If the feedback is positive then equation 2.13 becomes
𝐶(𝑠) 𝐺(𝑠)
= (2.14)
𝑅(𝑠) 1−𝐺(𝑠)𝐻(𝑠)

From equation (2.10) put the value of C(s) in equation (2.13)


𝐺(𝑠)𝐸(𝑠) 𝐺(𝑠)
=
𝑅(𝑠) 1+𝐺(𝑠)𝐻(𝑠)

𝐸(𝑠) 1
= (2.15)
𝑅(𝑠) 1+𝐺(𝑠)𝐻(𝑠)

Equation 2.15 represents the error ratio

𝐵(𝑠) 𝐺(𝑠)𝐻(𝑠)
Primary feedback ratio = = (2.16)
𝑅(𝑠) 1+𝐺(𝑠)𝐻(𝑠)

For positive feedback


𝐵(𝑠) 𝐺(𝑠)𝐻(𝑠)
=
𝑅(𝑠) 1−𝐺(𝑠)𝐻(𝑠)

2.5.1 Unity Feedback Control System

R(s) + C (s)
G(s)
-
B (s)

Figure 2.5: Unity feedback

For unity feedback control system, H(s) = 1

𝐶(𝑠) 𝐺(𝑠)
=
𝑅(𝑠) 1+𝐺(𝑠)

36
2.6 Cascaded Transfer Functions

2.6.1 Block Diagram Reduction


When a number of blocks are connected, the overall transfer function can be obtained by block
diagram reduction technique. The following are some of the rules associated with the technique.

Rule 1: Blocks in Cascade


When two or more blocks are in cascade, the resultant block is a product of the individual block
transfer function. Consider Figure 2.6.

R(s) C1(s) C(s)


G1(s) G2(s)

Figure 2.6: Blocks in Cascade

From figure 2.6,

𝐶1 (𝑠) 𝐶(𝑠)
= 𝐺1 (𝑠) and = 𝐺2 (𝑠)
𝑅(𝑠) 𝐶1 (𝑠)

From the above equations

𝐶1 (𝑠) 𝐶(𝑠) 𝐶(𝑠)


. = = 𝐺1 (𝑠). 𝐺2 (𝑠)
𝑅(𝑠) 𝐶1 (𝑠) 𝑅(𝑠)

The equivalent diagram is shown in Figure 2.6b

R(s) C(s)
G1(s).G2(s)

Figure 2.6b: Resultant

Rule 2: Blocks in Parallel


When 2 or more blocks are connected in parallel, the resultant block is the sum of individual block
transfer function.

G1(s)
+
+
R(s) G2(s) C(s)
+
G3(s)

Figure 2.7: Blocks in Parallel

𝐶(𝑠) = 𝑅(𝑠)𝐺1 (𝑠) + 𝑅(𝑠)𝐺2 (𝑠) + 𝑅(𝑠)𝐺3 (𝑠)


= 𝑅(𝑆)[𝐺1 (𝑠) + 𝐺2 (𝑠) + 𝐺3 (𝑠)]
𝐶(𝑠)
= 𝐺1 (𝑠) + 𝐺2 (𝑠) + 𝐺3 (𝑠)
𝑅(𝑠)

37
Rule 3: Moving a takeoff point ahead of a Block
If a takeoff point is moved ahead of a block, a block with same transfer function is introduced in the
branch of the takeoff point.

R(s) R(s) G(s) C(s)


G(s) C(s)


G(s)
Figure 2.8: Take-off point ahead of block

Rule 4: Moving a takeoff point after a Block


If a takeoff point is moved after a block, a block with a reciprocal of the transfer function is introduced
in the branch of a takeoff point as shown in figure 2.9.
R(s) C(s)
R(s) G(s) G(s)
C(s)
≡ 1
𝐺(𝑠)
Figure 2.9: Take-off point after a block

Rule 5: Moving a summing point beyond a Block


+ C (s) + C (s)
x G(s)
x G(s)
- ≡
y G(s)
y
𝐶(𝑠) = 𝐺(𝑠). (𝑥 − 𝑦) 𝐶(𝑠) = 𝐺(𝑠)𝑥 − 𝐺(𝑠)𝑦
= 𝐺(𝑠)(𝑥 − 𝑦)
Figure 2.10: Summing beyond a block

Rule 6: Moving a summing point ahead of a block

R1(s) C(s) R1(s) + C(s)


G(s) G(s)
± ±
≡ 1
R2(s)
R2(s) G(s)

𝑅2 (𝑆)
𝐶(𝑠) = 𝑅1 (𝑆)𝐺(𝑆) ± 𝑅2 (𝑆) 𝐶(𝑠) = [𝑅1 (𝑆) ± ] 𝐺(𝑆)
𝐺(𝑆)

= 𝑅1 (𝑆)𝐺(𝑆) ± 𝑅2 (𝑆)

Figure 2.11: Summing point moved before block

38
2.6.2 Open loop System
This sub-topic describes how an open loop control system can be reduced based on the rules
aforementioned.

Example 2.4
Derive the transfer function of Figure 2.12a using block reduction technique.

G4

M1
R G1 G2 G3 C

(a)

Solution 2.4
Step 1:
The two blocks, G1 and G2 are in cascade, so rule 1 must be applied.

G4

R G1 G2 G3 C

(b)

Step 2:
But G3 and G4 are parallel, hence, rule 2 is applied.

M1
R G1 G2 G3 + G4 C

(c)

Step 3:
Figure 2.12 c represents two blocks in cascade so rule 1 is applied again.

R G1 G2 (G3 + G4) C

(d)

Figure 2.12: Block-diagram reduction for an open-loop system

𝐶
Thus = 𝐺1 𝐺2 (𝐺3 + 𝐺4 )
𝑅

39
2.6.3 Closed loop system
Example 2.5
Find the overall transfer function of the system shown in figure 2.13.

G4

R + E M + M1 + C
G1 G2 G3 + G5
-
- -

B B1 B2
H1 H2

(a)

R + E G1 G2 G5 C
G3+G4
- 1+G2H1 1+G5H2
B

(b)

R + G1 G2 (G3+G4) G5
C
- (1+G2H1)(1+G5H2)

(c)
Figure 2.13: Determining the overall transfer function for closed-loop
system

From figure 2.13 (c), the system is a closed-loop system but the 𝐻 = 1 (unity feedback),
𝐶 𝐺
thus =
𝑅 1+𝐺

𝐺1 𝐺2 (𝐺3 + 𝐺4 )𝐺5
𝐶 (1 + 𝐺2 𝐻1 )(1 + 𝐺5 𝐻2 )
=
𝑅 𝐺1 𝐺2 (𝐺3 + 𝐺4 )𝐺5
1+
(1 + 𝐺2 𝐻1 )(1 + 𝐺5 𝐻2)

𝐶 𝐺1 𝐺2 (𝐺3 + 𝐺4 )𝐺5
=
𝑅 (1 + 𝐺2 𝐻1 )(1 + 𝐺5 𝐻2 ) + 𝐺1 𝐺2 (𝐺3 + 𝐺4 )𝐺5

40
Example 2.6
𝐶(𝑠)
Determine the ratio for the system shown in figure 2.14
𝑅(𝑠)

H2

-
R(s) + + G1 + G2 G3 C(s)
- -

H1

Figure 2.14: Diagram for example 2.6

Solution 2.6

Step 1: Shift the takeoff point beyond the block G3


H2

-
R(s) + + G1 + G2 G3 C(s)
- -

𝐻1
𝐺3

a: For step 1

Step 2: G2 and G3 are in cascade

H2

-
R(s) + + G1 + G2G3 C(s)
- -

𝐻1
𝐺3

b: For step 2

41
R(s) + + 𝐺2 𝐺3
G1
1 + 𝐺2 𝐺3 𝐻2 C(s)
- -

𝐻1
𝐺3

c: Step 3

R(s) + + 𝐺1 𝐺2 𝐺3 C(s)
1 + 𝐺2 𝐺3 𝐻2
- -

𝐻1
𝐺3

d: Step 4

R(s) + 𝐺1 𝐺2 𝐺3
1 + 𝐺2 𝐺3 𝐻2 + 𝐺1 𝐺2 𝐻1 C(s)
-

e: Step 5

R(s) 𝐺1 𝐺2 𝐺3
C(s)
1 + 𝐺2 𝐺3 𝐻2 + 𝐺1 𝐺2 𝐻1 + 𝐺1 𝐺2 𝐺3

f: Step 6

Figure 2.14: Block reduction for example 2.6

𝐶(𝑠) 𝐺1 𝐺2 𝐺3
=
𝑅(𝑠) 1 + 𝐺2 𝐺3 𝐻2 + 𝐺1 𝐺2 𝐻1 + 𝐺1 𝐺2 𝐺3

42
2.7 Stability of Linear Systems
The concept of stability is very important to analyse and design the system. Stability could be
described in several ways.
i. A system is stable if the natural response approaches zero for sufficiently large time.
ii. A system is unstable if the natural response approaches infinity as time approaches infinity.
iii. A system is marginally stable if the natural response neither decays nor grows but approaches a
constant value for a sufficiently large time.
Physically, an unstable system whose natural response grows without bound can cause damage to the
system, to adjacent property or to human life. Many times, systems are designed with limit stops to
prevent total runaway. From the perspective of the time response plot of a physical system, instability
is displayed by transients that grow without bound and, consequently, a total response that does not
approach a steady-state value or other forced response.

2.8 How to Determine a Stable System

Figure 2.15 Closed-loop Poles and Response: (a) Stable system; (b) Unstable
System
A study of system poles while considering natural response definitions of stability reveal that poles
in the left-half plane (LHP) yield either pure exponential decay or damped sinusoidal natural
responses. These natural responses decay to zero as time approaches infinity. Thus, if the closed-loop
system poles are in the left half of the s-plane and hence have a negative real part, the system is stable.
That is, stable systems have closed-loop transfer functions with poles only on the left half-plane.

43
Poles in the right half-plane (RHP) yield either pure exponentially increasing or exponentially
increasing sinusoidal natural responses. These natural responses approach infinity as time approaches
infinity. Thus, if the closed-loop system poles are in the right half of the s-plane and hence have a
positive real part, the system is unstable. Also, poles of multiplicity greater than one on the imaginary
axis lead to the sum of responses of the form 𝐴𝑡 𝑛 cos(𝜔𝑡 + 𝜑),
where 𝑛 = 1,2, …, which also approaches infinity as time approaches infinity. Thus unstable systems
have closed-loop transfer functions with at least one pole in the right half-plane and/or poles of
multiplicity greater than one on the imaginary axis.
Finally a system that has imaginary axis poles of multiplicity 1 yields pure sinusoidal oscillations as
a natural response. These responses neither increase nor decrease in amplitude. Thus marginally
stable systems have closed-loop transfer functions with only imaginary axis poles of multiplicity 1
and poles in the left half-plane.
For example, the unit step response of the stable system of Figure 2.15a, as compared to the unstable
system of Figure 2.15b, show that while the oscillations for the stable system diminish, those for the
unstable system increase without bound. Notice also that the stable system’s response in this case
approaches a steady-state value of unity.
It is not always a simple matter to determine if a feedback control system is stable even if we know
the poles of the corresponding forward transfer function as in Figure 2.16.

Figure 2.16 Common Cause of Problems in Finding Closed-loop Poles: a) Original System; b)
Equivalent System

However, under certain conditions, some conclusions may be drawn about the stability of the system.
First, if the closed-loop transfer function has only left half-plane poles, then the factors of the
denominator of the closed-loop system transfer function consist of products of terms such as (𝑠 + 𝑎𝑖 ),
where 𝑎𝑖 is real and positive, or complex with a positive real part. The product of such terms is a
polynomial with all positive coefficients. No term of the polynomial can be missing, since that would
imply cancellation between positive and negative coefficients or imaginary axis roots in the factors,
which is not the case.
Thus, a sufficient condition for a system to be unstable is that all signs of the coefficients of the
denominator of the closed-loop transfer function are not the same. If powers of s are missing, the
system is either unstable or, at best, marginally stable. Unfortunately, if all coefficients of the
denominator are positive and not missing, we do not have definitive information about the system’s
pole locations.

2.9 Routh-Hurwitz Criterion


The Routh-Hurwitz stability criterion provides a simple algorithm to decide whether or not the
zeros of a polynomial are all in the left half of the complex plane (such a polynomial is called at
times "Hurwitz"). A Hurwitz polynomial is a key requirement for a linear continuous-time time
invariant to be stable;
44
i. Necessary stability conditions
They are conditions that must hold for a polynomial to be Hurwitz. If any of them fails - the
polynomial is not stable. However, they may all hold without implying stability.

ii. Sufficient stability conditions


Conditions that if met imply that the polynomial is stable. However, a polynomial may be stable
without implying some or any of them.

The Routh criteria provide conditions that are both necessary and sufficient for a polynomial to be
Hurwitz. The Routh-Hurwitz criterion is comprised of three separate tests that must be satisfied. If
any single test fails, the system is not stable and further tests need not be performed. For this
reason, the tests are arranged in order from the easiest to determine to the hardest.

The Routh Hurwitz test is performed on the denominator of the transfer function, the characteristic
equation. For instance, in a closed-loop transfer function with G(s) in the forward path, and H(s) in
the feedback loop, we have:
𝐺(𝑠)
𝑇(𝑠) =
1+𝐺(𝑠)𝐻(𝑠)

If we simplify this equation, we will have an equation with a numerator P(s), and a denominator
Q(s):
𝑃(𝑠)
𝑇(𝑠) =
𝑄(𝑠)
The Routh-Hurwitz criteria will focus on the denominator polynomial Q(s).

2.10 Routh-Hurwitz Tests


Here are the three tests of the Routh-Hurwitz Criteria. For convenience, we will use n as the order
of the polynomial (the value of the highest exponent of s in Q(s)). The equation Q(s) can be
represented generally as follows:
𝑄(𝑠) = 𝑎𝑛 𝑠 𝑛 + 𝑎𝑛−1 𝑠 𝑛−1 + 𝑎𝑛−2 𝑠 𝑛−2 + ⋯ + 𝑎2 𝑠 2 + 𝑎1 𝑠1 + 𝑎0 𝑠 0

Rule 1: All the coefficients 𝑎𝑖 must be present (non-zero). That is, there should be no missing term.

Rule 2: All the coefficients of the equation should have the same sign. That is 𝑎𝑖 must be positive
(equivalently all of them must be negative, with no sign change)

Rule 3: If Rule 1 and Rule 2 are both satisfied, then form a Routh array from the coefficients 𝑎𝑖 .
There is one pole in the right-hand s-plane for every sign change of the pivot elements in the
Routh array (any sign change, therefore, means the system is unstable).

NB:
If the above two conditions are not satisfied the system will be unstable. But if all the coefficients
have the same sign and there is no missing term, there is no guarantee that the system will be stable.
In this section we learn a method that gives stability information without actually solving for the
closed-loop system poles. Using this method, we can tell how many closed-loop system poles are in
the left half-plane (LHP), right half-plane (RHP), and on the jω-axis. (Notice that we say how many,
not where). We can find the number of poles in each section of the s-plane, but not their co-ordinates
(Routh, 1905).

45
2.10.1 Statement of Routh-Hurwitz Criterion
Routh Hurwitz criterion states that the system is stable if and only if all the elements in the first
column have the same algebraic sign.
This method requires two steps:
i. Generate a data array called a Routh array and
ii. Interpret the Routh array to tell how many closed-loop system poles are in the LHP, RHP
and on the jω-axis.

2.10.2 Generating a Basic Routh Array


For the characteristic equation
𝑄(𝑠) = 1 + 𝐺(𝑠)𝐻(𝑠) = 0
Where
𝑄(𝑠) = 𝑎𝑛 𝑠 𝑛 + 𝑎𝑛−1 𝑠 𝑛−1 + 𝑎𝑛−2 𝑠 𝑛−2 + ⋯ + 𝑎2 𝑠 2 + 𝑎1 𝑠1 + 𝑎0 𝑠 0
The Routh array is formed by taking all the coefficients ai of Q(s), and staggering them in array
form.
Sn an an-2 an-4 ….. 0
S n-1
an-1
an-3
an-5
….. 0

That is, begin by labelling the rows with powers of s from the highest power of the denominator of
the closed-loop transfer function to 𝑠 0 . Next, start with the coefficient of the highest power of s in
the denominator and list, horizontally in the first row, every other coefficient. In the second row, list
horizontally starting with the next highest power of s, every coefficient that was skipped in the first
row. The final columns for each row should contain zeros:

Therefore, if n is odd, the top row will be all the odd coefficients. If n is even, the top row will be
all the even coefficients. We can fill in the remainder of the Routh Array as follows:
Sn an an-2 an-4 0
Sn-1 an-1 an-3 an-5 0
Sn-2 b1 b2 b3
. . .
. . .
S2 h1 h2
S1 j1 j2
S0 k1
−1 𝑎𝑛 𝑎𝑛−2 𝑎𝑛−1 𝑎𝑛−2 − 𝑎𝑛 𝑎𝑛−3
Where 𝑏1 = |𝑎 𝑎𝑛−3 | or 𝑏1 =
𝑎𝑛−1 𝑛−1 𝑎𝑛−1

and
−1 𝑎𝑛 𝑎𝑛−4 𝑎𝑛−1 𝑎𝑛−4 − 𝑎𝑛 𝑎𝑛−5
𝑏2 = |𝑎 𝑎𝑛−5 | or 𝑏2 =
𝑎𝑛−1 𝑛−1 𝑎𝑛−1

For each row that we are computing, we call the left-most element in the row directly above it the
pivot element. For instance, in row b, the pivot element is an-1, and in row c, the pivot element is b1
and so on, until we reach the bottom of the array.
Each entry is a negative determinant of entries in the previous two rows divided by the pivot element
the calculated row. The left-hand column of the determinant is always the first column of the previous
two rows, and the right-hand column is the elements of column above and to the right. The table is
complete when all of the rows are completed down to 𝑠 0 .

46
Example 2.7: Stable Third Order System

We are given a system with the following characteristic equation:


Q(s) = s3 + 2s2 + 4s + 3

Using the first two requirements, we see that all the coefficients are non-zero, and all of the
coefficients are positive. We will proceed then to construct the Routh-Array:

And we can calculate out all the coefficients:

And filling these values into our Routh Array, we can determine whether the system is stable:

From this array, we can clearly see that all of the signs of the first column are positive, there are no
sign changes, and therefore there are no poles of the characteristic equation in the RHP.

Example 2.8: Unstable Fifth Order System

Consider a system with a characteristic polynomial


𝑄(𝑠) = 𝑠 5 + 4𝑠 4 + 2𝑠 3 + 5𝑠 2 + 3𝑠 + 6
We have the following:

S5 1 2 3 0
S4 4 5 6 0
3 6
S3 3
/4 3
/2 0 0
2
S -3 6 0
S1 12 0
S0 6 0
In the first column, there are two sign changes (3 → −3, and −3 → 12), thus there are two non-
negative poles and the system is unstable.

47
2.11 Special Cases of Routh-Hurwitz Criterion
Two special cases can occur:
i) The Routh table may have an entire row consisting of zeros or
ii) It will sometimes have zero in the first column of a row.

2.11.1 Special Case 1: Row of All Zeros


If, while calculating our Routh-Hurwitz, we obtain a row of all zeros, we do not stop, but can
actually learn more information about our system. When any one row of Routh table is zero, it
indicates that the equation has at least one pair of roots which lie radially opposite each other and
equidistant from origin.
The array can be completed by forming the auxiliary polynomial. This is the polynomial whose
coefficients are the elements of the row just above the row of zeros in the Routh array. The
auxiliary polynomial can be very helpful.
The roots of the auxiliary polynomial give us the precise locations of complex conjugate roots that
lie on the 𝑗𝜔 axis. However, one important point to notice is that if there are repeated roots on the
𝑗𝜔 axis, the system is actually unstable. Therefore, we must use the auxiliary polynomial to
determine whether the roots are repeated or not.

Example 2.9
Determine the number of RHP poles and the stability of the closed-loop transfer function;
𝑄(𝑠) = 1 + 𝐺(𝑆)𝐻(𝑠) = 𝑠 5 + 7𝑠 4 + 6𝑠 3 + 42𝑠 2 + 8𝑠 + 56

Solution 2.9:
Start by forming the Routh table for the denominator as shown in table. At the second row, divide
through by 7 for convenience since 7 is a factor of all the elements in that particular row.
At the third row, the entire row consists of zeros.
Routh array for entire row is zero
5
s 1 6 8
s4 7 1 42 6 56 8
3
s 0 4 1 0 12 3 0 0 0
2
s 3 8 0
s1 1
/3 0 0
0
s 8 0 0

Now use the following procedure to proceed:


First, return to the row just above the row of zeros (𝑆 4 in this case) and form an auxiliary polynomial
using the entries in that row as coefficients. The polynomial will start with the power of s in the label
column and continue by skipping every other power of s. Thus the polynomial formed is
𝑃(𝑠) = 𝑠 4 + 6𝑠 2 + 8
Next, differentiate the polynomial with respect to 𝑠 and obtain;
𝑑𝑃(𝑠)
= 4𝑠 3 + 12𝑠 + 0
𝑑𝑠
Finally, use the coefficients of the resulting differentiation to replace the row of zeros. Again, for
convenience, the third row is divided by 4 after replacing the zeros. The array is completed afterwards
following standard procedures.
From the results, there is no pole in the right half plane, since there are no sign changes.
48
However, the system may or may not be stable, hence there is the need to proceed further to
determine the pair of roots and their exact locations. This is obtained from the roots of the auxiliary
polynomial, 𝑃(𝑠) = 𝑠 4 + 6𝑠 2 + 8 = 0.
Now from 𝑃(𝑠) = 𝑠 4 + 6𝑠 2 + 8 = 0
That breaks down to (𝑠 2 + 4)(𝑠 2 + 2) = 0
hence 𝑠 = ±𝑗2 𝑜𝑟 ± 𝑗√2
To get the other factor of Q, then 𝑠 5 + 7𝑠 4 + 6𝑠 3 + 42𝑠 2 + 8𝑠 + 56 must divide 𝑠 4 + 6𝑠 2 + 8
The resultant is 𝑠 + 7
Hence the factors of s are (𝑠 + 7)(𝑠 + 𝑗2)(𝑠 − 𝑗2)(𝑠 + 𝑗√2)(𝑠 − 𝑗√2)

Digression
To prove that the auxiliary polynomial 𝑃(𝑠) is a factor of 𝑄(𝑠), determine 𝑠 for 𝑃(𝑠)
Now putting say, 𝑠 = 𝑗2 𝑖𝑛 𝑄(𝑠) = 𝑠 5 + 7𝑠 4 + 6𝑠 3 + 42𝑠 2 + 8𝑠 + 56
Then 𝑄(𝑗2) = (𝑗2)5 + 7(𝑗2)4 + 6(𝑗2)3 + 42(𝑗2)2 + 8(𝑗2) + 56
𝑄(𝑗2) = 𝑗32 + 112 − 𝑗48 − 168 + 𝑗16 + 56
𝑄(𝑗2) = 0

Example 2.10
Apply Routh-Hurwitz criterion to the equation 𝑠 6 + 2𝑠 5 + 8𝑠 4 + 12𝑠 3 + 20𝑠 2 + 16𝑠 + 16 = 0
and investigate its stability, stating the number of roots with
(i) positive real part (ii) zero real part (iii) negative real parts

Solution 2.10
𝑄(𝑠) = 𝑠 6 + 2𝑠 5 + 8𝑠 4 + 12𝑠 3 + 20𝑠 2 + 16𝑠 + 16 = 0

𝑠6 1 8 20 16
𝑠5 2 1 12 6 16 8 0
𝑠4 2 1 12 6 16 8
𝑠3 0 4 1 0 12 3 0
𝑠2 6 3 16 8
𝑠1 1⁄ 0
3
𝑠0 8 0

Auxiliary polynomial, 𝑃(𝑠) = 𝑠 4 + 6𝑠 2 + 8


𝑑𝑃
= 4𝑠 3 + 12𝑠
𝑑𝑆
From the table, the system may be stable.
𝑠 4 + 6𝑠 2 + 8 = 0
Let 𝑥 = 𝑠 2
Hence
𝑥 2 + 6𝑥 + 8 = 0 or (𝑥 + 2)(𝑥 + 4) = 0
𝑥 = −2 𝑜𝑟 𝑥 = −4
⟹ 𝑠 2 = −2 𝑜𝑟 𝑠 2 = −4 Or 𝑠 = ±𝑗√2 𝑜𝑟 𝑠 = ±𝑗2

49
Now dividing 𝑄(𝑠) by 𝑃(𝑠)
𝑠6 +2𝑠5 +8𝑠4 +12𝑠3 +20𝑠2 +16𝑠+16
𝑠4 +6𝑠2 +8
This results in 𝑠 2 + 2𝑠 + 2 or 𝑠 = −1 ± 𝑗

Thus overall
𝑠1 = +𝑗√2; 𝑠2 = −𝑗√2; 𝑠3 = +𝑗2; 𝑠4 = −𝑗2; 𝑠5 = −1 + 𝑗; 𝑠6 = −1 − 𝑗

Hence
(i) There are no roots with positive real part
(ii) There are four (4) roots with zero real part
(iii) There are two (2) roots with negative real part

2.11.2 Special Case 2: Zero in the First Column (Pivot Element is zero)
In this special case, there is a zero in the first column of the Routh Array, but the other elements of
that row are non-zero; division by zero would be required to form the next row. Like the above case,
we can replace the zero with a value ε that we define as being an infinitely small positive number and
use that variable to continue our calculations. After the entire array has been constructed we can take
the limit as epsilon approaches 0 from either the positive or the negative side to determine the final
format of our Routh array.
Another approach is by defining a polynomial with reverse coefficients to the original polynomial.
Yet, a third approach is to multiply the original equation by a factor (𝑠 + 𝑎).

Using Epsilon Method


Example 2.11
Determine whether 𝑄(𝑠) = 1 + 𝐺(𝑆)𝐻(𝑆) = 𝑠 5 + 𝑠 4 + 4𝑠 3 + 4𝑠 2 + 2𝑠 + 1 = 0 is stable or not.

Solution 2.11: Replace the zero by a small positive quantity 𝜀 and continue the procedure.

S5 1 4 2
S4 1 4 1
S3 ɛ 1 0
0
S2 4𝜀−1 1 0
𝜀
4𝜀−1
𝜀
−𝜀 0
S1 4𝜀−1
𝜀
𝜀2
=1−
4𝜀 − 1
S0 1

50
Now checking
If 𝜺 𝒊𝒔 + 𝒗𝒆
𝑠 2 = −𝑣𝑒; 𝑠1 = +𝑣𝑒 and 𝑠 0 = 1
Hence 𝑠 5 = +𝑣𝑒; 𝑠 4 = +𝑣𝑒; 𝑠 3 = +𝑣𝑒; 𝑠 2 = −𝑣𝑒; 𝑠1 = +𝑣𝑒; 𝑠 0 = +𝑣𝑒
There are two changes +→ −→ +, thus system is unstable and has two RHP poles.
.
If 𝜺 𝒊𝒔 – 𝒗𝒆
𝑠 2 = +𝑣𝑒; 𝑠1 = +𝑣𝑒 and 𝑠 0 = 1
Hence 𝑠 5 = +𝑣𝑒; 𝑠 4 = +𝑣𝑒; 𝑠 3 = −𝑣𝑒; 𝑠 2 = +𝑣𝑒; 𝑠1 = +𝑣𝑒; 𝑠 0 = +𝑣𝑒
There are two changes +→ −→ +, thus system is unstable and has two RHP poles.

Reverse Coefficients Method


A polynomial that has the reciprocal roots of the original polynomial has its roots distributed the
same on the s-plane because taking the reciprocal of the root plane does not move the poles to another
region. Thus, if polynomial that has the reciprocal roots of the original could be found, it is possible
that the Routh table for the new polynomial will not have any zero in the 1st column.
This polynomial is simply the original polynomial with its coefficients written in reverse order
(Phillips, 1991).

Example 2.12:
Determine the stability of the closed-loop transfer function;
T (s ) =
10
s + 2s + 3s 3 + 6s 2 + 5s + 3
5 4

Solution: First, test the polynomial by applying it direct through the Routh table to see whether it has
any of the special cases mentioned.
s5 1 3 5
s4 2 6 3
s3 0 3.5

Since this polynomial (i.e. the denominator) gives a zero in the first column of the Routh table when
applied direct, write the reciprocal roots of the denominator in reverse order as shown below;
D(s ) = 3s 5 + 5s 4 + 6s 3 + 3s 2 + 2s + 1
The polynomial D(s) is the polynomial with the reverse coefficients to the original polynomial in
T(s). Using the reverse polynomial, the Routh table is now formed as shown in table.

s5 3 6 2
4
s 5 3 1
s3 4.2 1.4
s2 1.33 1
s1 -1.75
s0 1

Since there are 2 sign changes, the system is unstable and has two RHP poles.

51
Using 𝒔 + 𝒂 method
In this method, the original equation is multiplied by a factor (𝑠 + 𝑎), where ‘a’ is any positive real
number. The simplest value of ‘a’ is 1 (take a = 1).

Example 2.13
Investigate the stability of
𝑄(𝑠) = 𝑠 5 + 𝑠 4 + 2𝑠 3 + 2𝑠 2 + 3𝑠 + 5 = 0

Solution 2.13

S5 1 2 3
4
S 1 2 5
3
S 0 -2
S2
S1
S0

Now multiply the characteristic equation by (s+1)


(𝑠 5 + 𝑠 4 + 2𝑠 3 + 2𝑠 2 + 3𝑠 + 5)(𝑠 + 1)
That is, 𝑠 6 + 2𝑠 5 + 3𝑠 4 + 4𝑠 3 + 5𝑠 2 + 8𝑠 + 5

S6 1 3 5 5
S5 2 4 8
S4 1 1 5
S3 2 -2
S2 2 5
S1 -7
S0 5

From the table:


Since there are two sign changes in the pivot column, then there are 2 poles in the RHP, hence system
is unstable.

Illustrating Pole Distribution for the Row of Zeros


Example 2.14
Tell how many poles are in the RHP, LHP and on the 𝑗𝜔 −axis for the closed-loop transfer function
given below.
𝑄(𝑠) = 𝑠 8 + 𝑠 7 + 12𝑠 6 + 22𝑠 5 + 39𝑠 4 + 59𝑠 3 + 48𝑠 2 + 38𝑠 + 20

52
Solution 2.14
The entire solution following the outlined procedures is shown below.

s8 1 12 39 48 20
s7 1 22 59 38 0
s6 -10 -1 -20 -2 10 1 20 2 0
s5 20 1 60 3 40 2 0 0
s4 1 3 2 0 0
s3 0 4 2 0 6 3 0 0 0 0 0
s2 3
/2 3 2 4 0 0 0
s1 1
/3 0 0 0 0
s0 4 0 0 0 0

The even polynomial appears in the row directly above the row of zeros. Every entry in the table
from the even polynomial’s row to the end of the chart applies only to the even polynomial.
Therefore, the number of sign changes from the even polynomial to the end of the table equals the
number of RHP roots of the even polynomial. Because of the symmetry of the roots about the origin,
the even polynomial must have the same number of LHP roots as it does RHP roots. Having
accounted for the roots in the right and LHPs, the remaining roots must be on the j -axis. Hence,
example 2.14 has the following summary of pole distributions shown in table below.

Table 2.1 Summary of Pole Locations for Example 2.14


Location Even (fourth-order) Other (fourth-order) Total (eighth-order)
RHP 0 2 2
LHP 0 2 2
j 4 0 4

One method of showing the stability of a system has been given by Routh–Hurwitz Criterion. The
main difficulty associated with this method is that the exact location or coordinates of the poles
cannot be given by Routh-Hurwitz. In any case, you would be able to tell whether the system is
stable, unstable or marginally stable (where all the roots are on the j -axis). Again, to handle this
method effectively and practically, one need to know how to derive transfer functions from both
electrical and electromechanical systems. This area together with the analysis and design of feedback
systems would be captured in a subsequent course offered by the department namely Control
Systems.

2.12 Problem

1. Draw the block diagram of a closed-loop control system and indicate the following on it:
(i) Plant (ii) command input (iii) controlled output
(iv) Actuating signal (v) feedback element and (vi) control element

2. An amplifier with a negative feedback provides an output voltage of 5 𝑉 with an input voltage
of 0.2 𝑉. On removing feedback it requires only 0.1 𝑉 input to provide the same output.
Calculate:
(i) Gain without feedback
(ii) Gain with feedback
(iii) Feedback ratio
[𝑨 = 𝟓𝟎; 𝑨𝒇 = 𝟐𝟓; 𝜷 = 𝟎. 𝟎𝟐]
53
3. The voltage gain of an amplifier without feedback is 2000. Calculate the voltage gain of the
amplifier if negative feedback is introduced in the circuit. Assume that feedback factor is
0.01. [𝑨𝒇 = 𝟗𝟓. 𝟐𝟒]

4. Use block diagram reduction method to obtain the equivalent transfer function from R to C.

H2

-
R(s) + + +
G3 G4
G1 G2 C(s)
- -

H3

H1

5. When negative feedback is applied to an amplifier of gain 200, the overall gain falls to 50.
(i) Calculate the value of feedback
(ii) If the fraction i.e. feedback factor remains the same, calculate the value of amplifier
gain so that the overall gain becomes 40.
[𝜷 = 𝟎. 𝟎𝟏𝟓; 𝑨 = 𝟏𝟎𝟎]

6. A closed loop control system has the characteristic equation given by


𝑠 3 + 4.5𝑠 2 + 3.5𝑠 + 1.5 = 0.
Investigate the stability using Routh-Hurwitz criterion.

7. Make a Routh table and tell how many roots of the following polynomial are in the RHP and
in the LHP.
𝑄(𝑠) = 3𝑠 7 + 9𝑠 6 + 6𝑠 5 + 4𝑠 4 + 7𝑠 3 + 8𝑠 2 + 2𝑠 + 6

8. Apply Routh-Hurwitz criterion to the following equation and investigate the stability
𝑠 5 + 2𝑠 4 + 2𝑠 3 + 4𝑠 2 + 11𝑠 + 10 = 0

54
CHAPTER 3
ACTIVE FILTERS
3.0 Introduction
A filter is a circuit that passes certain frequencies or frequency ranges and attenuates or rejects
other frequencies. In other words, filters are capable of passing input signals with certain selected
frequencies through to the output while rejecting signals with other frequencies. This property is
called selectivity. A filter circuit, therefore, possesses at least one passband – a band of
frequencies in which the output is approximately equal to the input (i.e. attenuation is zero) and an
attenuation band in which the output is zero (i.e. attenuation is infinite).
As frequency selected devices, they are used in radio and TV receivers to allow one to select one
desired signal out of a multitude of broadcast signals in the environment.

3.1 Naming of Filters


Filters may be of any type such as electrical, mechanical, pneumatic, hydraulic, acoustical etc., but
the most commonly used filters are of the electric type. They are variously named by
• What they pass or suppress e.g. low-pass, band-pass, high-pass, band reject (stop) or notch.
• Part of frequency spectrum where they are used e.g. audio, microwave, infrared, ultra-violet,
x-ray, etc.
• Bandwidths e.g. wideband, narrowband.
• many other ways

3.1.1 Electrical Filters


They are used in practically all circuits which require separation of signals according to their
frequencies. Applications include (but are not limited to) noise rejection and signal separation in
industrial and measurement circuits, feedback of phase and amplitude control in servo-loops,
smoothing of digitally generated analog signals, audio-signal shaping and sound enhancement,
channel separation, and signal enhancement in communication circuits.

Application of Electrical Filters


• Separation of direct current from a mixture of direct and alternating current
• Acquiring alternating current from a mixture of direct and alternating current
• Selecting one or more desired radio signals from one or more undesired signals and noise
• Obtaining upper- or lower-sidebands from double-sideband (DSB) signals
• carrier-waves or sidebands from amplitude-modulated (AM) signals
• long pulses or short pulses from mixtures of long and short pulses

3.2 Active and Passive Filters


3.2.1 Passive filter
A filter is a passive filter if it consists of only passive elements Resistors, R, Inductors, L and
Capacitors, C (that is, RC, RL, LC or RLC). They are also called passive filters because they do not
depend on an external power supply and/or do not contain active components such as transistors
They are most responsive to frequencies between around 100 Hz and 300 MHz. The lower frequency
limit results from the fact that at low frequencies the capacitance and inductance values become
exceedingly large, meaning prohibitively large components are needed.
When designing passive filters with very steep attenuation/falloff responses, the number of inductor
and capacitor sections must increase.
55
Advantages
i. no, amplifying elements (transistors, op-amps, etc.)
ii. no signal gain
iii. 1st order - design is simple (just use standard equations to find resonant frequency of the
circuit)
iv. 2nd order - complex equations
v. require no power supplies
vi. not restricted by the bandwidth limitations of the op-amps
vii. can be used at very high frequencies
viii. can handle larger current or voltage levels than active devices

Disadvantages
i. high accuracy (1% or 2%), small physical size, or large inductance values are required
ii. standard values of inductors are not very closely spaced
iii. difficult to find an off-the-shelf inductor within 10 percent of any arbitrary value
iv. adjustable inductors are used
v. tuning such inductors to the required values is time-consuming and expensive for larger
quantities of filters
vi. inductors are often prohibitively expensive

3.2.2 Active Filter


A filter is said to be an active filter if it consists of active elements such as transistors and op-amps,
combined with passive elements.
• no inductors
• made up of op-amps, resistors and capacitors
• provides virtually any arbitrary gain
• generally easier to design
• high input impedance prevents excessive loading of the driving source
• low output impedance prevents the filter from being affected by the load
• at high frequencies is limited by the gain-bandwidth of the op-amps
• easy to adjust over a wide frequency range without altering the desired response

3.3 Terminologies Associated with Filters


3.3.1 -3-dB Frequency (f3dB).
This represents the input frequency that causes the output signal to drop to –3 dB relative to the input
signal. The –3 dB frequency is equivalent to the cutoff frequency; the point where the input-to-output
1
power is reduced by one-half (50%) or the point where the input-to-output voltage is reduced by
√2
(70.7%). For low-pass and high-pass filters, there is only one –3-dB frequency. However, for band-
pass and notch filters, there are two –3-dB frequencies, typically referred to as f1 and f2.

3.3.2 Passband
The pass band of a filter is the region of frequencies that are allowed to pass through the filter with
minimum attenuation, usually defined as less than – 3 decibels (dB) of attenuation. Ideally, the
attenuation is zero for the band of frequencies. Every filter has to have at least one pass band and at
least one attenuation band. For the attenuation band, ideally the attenuation is infinite for the band of
frequencies.

56
Av(dB)

0 dB
-3dB {

Stop-band
Passband

f
𝑓f𝑐2

Figure 3.1: A filter depicting – 3 dB frequency, passband and cut-off frequency

3.3.3 Critical Frequency, 𝑓𝑐


This is often called the cut-off frequency, 𝑓𝑐 defines the end of the passband and is normally specified
at the point where the response drops – 3 dB from the passband response.

3.3.4 Stop-band frequency (fs).


This is a specific frequency where the attenuation reaches a specified value set by the designer. For
low-pass and high-pass filters, the frequencies beyond the stop-band frequency are referred to as the
stop-band. For band-pass and notch filters, there are two stop-band frequencies, and the frequencies
between the stop-bands are also collectively called the stop-band frequencies.

3.3.5 Bandwidth
The bandwidth of a filter is a measure of its passband and is defined as the difference between the
upper and lower 3-dB cut-off frequencies of the passband.
𝐵𝑤 = 𝑓𝑐2 − 𝑓𝑐1
.

3.3.6 Roll-off Rate


This is the rate at which a gain drops, e.g., - 20 dB/decade. This means that at a frequency of 10 𝑓𝑐 ,
the output will be – 20 dB (10%) of the input.

3.3.7 Octave and Decade


To describe the rate of rise or fall in attenuation as the frequency changes, descriptions such as “6
dB/octave” or “20 dB/decade” are used.
𝜔
|𝐻(𝑗𝜔)| ≃ −20 log10
𝜔𝑜

𝜔 2
For octave apart, =
𝜔𝑜 1
⟹ |𝐻(𝑗𝜔)| ≃ −6 𝑑𝐵
Here, 6 dB/octave simply means that if the frequency changes by a factor of 2 (an octave), the
attenuation changes by 6 dB.
57
A decade is a ten times change in frequency.
𝜔 10
For decade apart, =
𝜔𝑜 1
⟹ |𝐻(𝑗𝜔)| ≃ −20 𝑑𝐵

20 dB/decade means that, if the frequency changes by a factor of 10, the attenuation increases or
decreases by 20 dB.
1
It must be noted that when |𝐻(𝑗𝜔)| ≃ 20 log10 , the half-power condition,
√2
then |𝐻(𝑗𝜔)| ≃ −3𝑑𝐵
This is called the –3-dB point. This point corresponds to the half-power point or cut-off frequency
for a filter.

NB
𝑉
❖ For voltages, gain G is 𝐺𝑑𝐵 = −20 log10 𝑉2
1
𝑃
❖ For Power, 𝐺𝑑𝐵 = −10 log10 𝑃2
1
𝐼
❖ For Current, 𝐺𝑑𝐵 = −20 log10 𝐼2
1

3.3.8 Pole and Order


A pole is nothing more than an RC circuit as could be observed in figure 3.2.
An n - pole filter contains n – RC circuits, typically between 1 and 6 for most applications.

+V
R1
+
vin
C1 vout
- Rf1
-V

Rf2

Order refers to the number of poles. That is,


1st order – 1 pole (single pole)
2nd order – 2 poles
nth order – n poles

58
3.3.9 Cascaded System

AV1 AV2 AV3


V1 X 10 X 10 X 10 Vo

-20 dB -20 dB -20 dB


Figure 3.3: A Cascaded System

𝐴𝑣 = 𝐴𝑣1 × 𝐴𝑣2 × 𝐴𝑣3


= 10 × 10 × 10
= 1000
Thus 𝐴𝑣 (𝑑𝐵) = −20 log10 1000
= −60𝑑𝐵

Alternatively,
𝐴𝑣 (𝑑𝐵) = 𝐴𝑣1 (𝑑𝐵) + 𝐴𝑣2 (𝑑𝐵) + 𝐴𝑣3 (𝑑𝐵)
= −20𝑑𝐵 + (−20𝑑𝐵) + (−20𝑑𝐵)
= −60𝑑𝐵

NB:
The number of poles determines the roll-off rate of the filter. For example, a Butterworth response
produces -20 dB/decade/pole. So a first order (one-pole) filter has a roll-off - 20 dB/decade, a
second-order (two-pole) filter – 40 dB/decade.

Table 3.1: Filter Orders and Roll-off Slopes


Order (Poles) dB/Octave dB/Decade Phase Shift * Comments
1st 6 20 90 Only passive, very common
2nd 12 40 180° Extremely common - most popular filter
3rd 18 60 270° Moderately common
4th 24 80 360° Linkwitz-Riley crossovers
5th 30 100 450° Very uncommon - rarely used
6th 36 120 540° Somewhat uncommon (ESP subsonic filter)
N n*6 n * 20 n * 90° Anti-aliasing filters (e.g. before ADC circuits)

❖ Phase shift refers to the phase difference between a high and low pass filter set for the same
roll-off frequency.

3.3.10 Critical Frequency and Roll-off Rate


The critical frequency is determined by the values of the resistors and capacitors in the RC circuit.
For a single – pole (first –order) filter, the critical frequency 𝑓𝑐 is
1
𝑓𝑐 =
2𝜋𝑅𝐶

59
3.4 Types of Filters
Ideally a filter would for example, completely eliminate signals above a cut-off frequency and
perfectly pass signals below it (in the pass-band). There is a wide range of filter circuits, each with
its own set of advantages and disadvantages. All filters introduce phase shift, and (almost all) filters
change the frequency response. In real filters, various trade-offs are made in an attempt to
approximate the ideal. Some filter types are optimized for gain flatness in the pass-band
(Butterworth), some trade off gain variation (ripple) in the pass-band for steeper roll-off (Chebyshev),
and still others trade off both flatness and rate of roll-off in favour of pulse-response fidelity (Bessel).
When describing how a filter behaves, a response curve is used, which is simply the attenuation or
gain (Vout/Vin) versus frequency graph (refer figure 3.4).

Gain

0 dB

fc f
Figure 3.4: An Ideal Low-pass (Brick wall) Filter

The types of filters to be discussed thoroughly in this chapter are:


a) low-pass
b) high-pass
c) band-pass
d) notch, or band-reject or band-stop

3.4.1 Low Pass Filter


This passes low frequencies and stops (attenuates) high frequencies. The low frequencies are from
dc (0 Hz) up to the upper cut-off frequency (𝑓𝐻 ). Thus frequencies higher than the cut-off frequency
are attenuated. A low-pass filter is sometimes called a high-cut filter, or treble cut filter when used
in audio applications.

Av(dB)

-3dB {
- 20 dB /decade

f
ffH2
`
Figure 3.5: Actual response of Low-pass Filter

60
The response drops gradually to zero at frequencies beyond the passband as depicted in figure 3.5.
This ideal response is sometimes referred to as “brick wall” because nothing gets through beyond the
wall. In the case of a low-pass filter the lower critical frequency is 0 Hz, so the bandwidth is equal to
𝑓𝐻 .

Bandwidth 𝐵𝑤 = 𝑓𝐻
1
𝑓𝐻 occurs when 𝑋𝑐 = 𝑅, 𝑤ℎ𝑒𝑟𝑒 𝑓𝐻 =
2𝜋𝑅𝐶
+V
R1
+
vin
C1 vout
- Rf1
-V

Rf2

Figure 3.6: Single-pole active low pass filter

The most basic low-pass filter is a simple RC circuit consisting of just one resistor and one capacitor.
This basic RC filter as observed in figure 3.6 has a single pole and it decreases at - 20 dB/decade
beyond the critical frequency.

Figures 3.7 and 3.8 also show two-pole and three-pole low pass filters respectively.
C2

+V
R2 R1
+
vin
C1 vout
- Rf1
-V

Rf2

Figure 3.7: Two-pole low pass filter

Stage 1 Stage 2
C2

+V
R2 R1
+ +V
vin R3
+
C1
- Rf1 C3 vout
-V - Rf3
-V
Rf2

Rf4

Figure 3.8: Three-pole low pass filter

61
3.4.2 High - Pass Filter
It is the filter that significantly attenuates or rejects all frequencies below the cut-off frequency, 𝑓𝐿
and passes all frequencies above it. By expansion the 𝑓𝐿 is the frequency at which the output is 70.7%
of the input (or – 3 dB). It is sometimes called a low-cut filter; the terms bass-cut filter or rumble
filter are also used in audio applications.
Av(dB)

-
3dB {

Stopband Passband

f
fL f ff1
L
(a) Ideal filter (b) Practical filter
Figure 3.9: High-pass filter response

Ideally, the passband of a high-pass filter is all frequencies above the critical frequency. The high-
frequency response of practical circuits is limited by the op-amp or other components that make up
the filter.
A simple RC circuit consisting of a single resistor and capacitor can be configured as a high-pass
filter by taking the output across the resistor (figure 3.10). As in the case of the low-pass filter, the
basic RC circuit has a roll-off rate of – 20 dB/decade. Also the critical frequency for the basic high-
1
pass filter occurs when 𝑋𝑐 = 𝑅, 𝑤ℎ𝑒𝑟𝑒 𝑓𝐿 =
2𝜋𝑅𝐶
+V
C1
+
.vin R1
vout
- Rf1
-V

Rf2

Figure 3.10: Single-pole active high- pass filter

R2

+V
C2 C1
+
vin
R1 vout
- Rf1
-V

Rf2

Figure 3.11: Two-pole high pass filter

62
3.4.3 Band - Pass Filter
A band-pass filter is a circuit which is designed to pass signals only in a certain band of frequencies
while attenuating all signals outside the band. In other words, the filter passes all signals lying within
a band between a lower-frequency limit and an upper-frequency limit and essentially rejects all other
frequencies that are outside this specific band. A generalized band-pass response curve is shown in
Ffigure 3.12.

𝑓0
Figure 3.12: Frequency Response of Band-pass filter

The parameters of importance in the band-pass filter are the high and low cut-off frequencies
(𝑓1 𝑎𝑛𝑑 𝑓2 ), the bandwidth BW, the centre frequency, 𝑓0 and the selectivity or Q.
C2 R4

+V
R2 R1 +V
+ C4 C3
vin +
C1 - R3
Rf1 - vout
-V
Rf3
-V
Rf2
Rf4
Stage 1 Stage 2
Two-pole low-pass Two-pole high-pass

Figure 3.13: Two-stage Band - pass Filter

Multiple – Feedback Band-Pass Filter


Another type of filter configuration shown in Figure 3.14, is a multiple-feedback band-pass filter.

C2
Rf

+V
R1 C1
-
vin R2
vout
+

-V

Figure 3.14: Multiple-feedback band - pass filter


63
This filter has an advantage over the cascaded band-pass filter in Figure 3.13 because it is simpler
and can achieve narrower bandwidths. There are two feedback paths in this configuration, one
through 𝐶2 and the other through 𝑅𝑓 . This is the reason why it is called multiple-feedback.

Centre frequency (f0)


On a standard log plot, band-pass filters are geometrically symmetrical around the filter’s resonant
frequency or centre frequency, provided the response is plotted on linear-log graph paper (the
logarithmic axis representing the frequency). On standard-log paper, the centre frequency, 𝑓𝑜 is
related to the –3-dB frequencies by the following expression:
𝑓𝑜 = √𝑓1 × 𝑓2 , i.e., the geometric mean of the cut-off frequencies.
For narrow-band band-pass filters, where the ratio of f2 to f1 is less than 1.1, the response shape
approaches arithmetic symmetry. Here, we can approximate f0 by taking the average of –3-dB
frequencies:
𝑓1 +𝑓2
𝑓𝑜 =
2

Quality Factor
There are basically two types of band-pass filters, that is, wide band-pass and narrow band-pass
filters. Unfortunately, there is no set dividing line between the two. However, a band-pass filter is
defined as a wide band-pass if its figure of merit or quality, Q is less than 10 while the band-pass
filters with 𝑄 > 10 are called the narrow band-pass filters.
The quality factor (Q) of a band-pass filter is the ratio of the centre frequency to the bandwidth.
𝑓𝑜
𝑄=
𝐵𝑤
The value of Q is an indication of the selectivity of a band-pass filter. The higher the value of Q, the
narrower the bandwidth and the better the selectivity for a given value of 𝑓𝑜 .

Example
A certain band-pass filter has a centre frequency of 6 kHz and a bandwidth of 500 Hz. Determine the
Q and classify the filter as narrow-band or wide-band.
Solution
𝑓𝑜 6000
𝑄= = = 12
𝐵𝑤 500
Because Q > 10, this is a narrow-band filter.

3.4.4 Band - Stop Filter


This is also known as the band-reject, or band-elimination filter. It attenuates a single band of
frequencies and allows those on either side to pass through.
The band-stop filter, just like band-pass filter has two cut-off frequencies. It will pass above or below
a particular range of frequencies whose cut-off frequencies are defined by the components used in
the circuit. Any frequency in between these cut-off frequencies is attenuated. So, it has two pass-
bands and one stop-band. This type of filter passes all frequencies from zero up to the lower cut-off
frequency. Then, it blocks all the frequencies between the lower and upper cut-off frequencies and
finally, passes all frequencies above the upper cut-off frequency. The characteristics of the band-stop
filter is shown in Figure 3.15.

64
Figure 3.15 Frequency Response of Band-stop Filter
response
Thus, for a band-stop filter, the stopband is all the frequencies between the lower and upper cut-off
frequencies. The frequencies below the lower cut-off frequency and above the upper cut-off
frequency are the passband. An ideal bandstop filter has infinite attenuation in the stopband, no
attenuation in the passband, and two vertical transitions.
The bandwidth is the band of frequencies between the 3 – dB points just as in the case of the band-
pass filter response. Thus, the band-stop filter can be thought of as the opposite to that of the band-
pass filter since the frequencies within a certain bandwidth are rejected, and frequencies outside the
bandwidth are passed.
The band stop filter’s circuit is made up of a high pass filter, a low-pass filter and a summing
amplifier. It is formed by combining the low pass and high pass filters in parallel connection through
an amplifier circuit. The summing amplifier will have an output that is equal to the sum of the filter
output voltages. The block diagram and the circuit diagram of the Band-stop filter are shown in
Figure 3.16 and Figure 3.17 respectively.

Figure 3.16 Block Diagram of the Band-stop Filter

Figure 3.17 Circuit Diagram of the Band-stop Filter


65
The notch filter as depicted in Figure 3.18, is a special type of band stop filter. It is designed to block
all frequencies that fall within its bandwidth.

Figure 3.18 Frequency Response of Notch Filter


response

3.5 Filter Response Characteristics


Each type of filter (low-pass, high-pass, band-pass or band-stop) can be tailored by circuit component
values to have either, a Butterworth, Chebyshev, or Bessel characteristic.
Av
Chebyshev

Bessel

Butterworth
Bessel Butterworth
Chebyshev

f
Each of these characteristics
Figure is identified
3.18: Comparative plotsby
of the
threeshape
typesofof the
filterresponse
responsecurve, and each has an
characteristics
advantage inresponse
certain applications. In this section, you will learn the three basic filter response
characteristics and other filter parameters.
Butterworth, Chebyshev, or Bessel response characteristics can be realized with most active filter
circuit configurations by proper selection of certain component values. A general comparison of the
three response characteristics for a low-pass filter response curve is shown in figure 3.16. The other
filters can also be designed to have any one of the characteristics.

3.5.1 The Butterworth Characteristic


The Butterworth characteristic provides a very flat amplitude response in the passband and a roll-off
rate of – 20 dB/decade/pole. The phase response is not linear, however, and the phase shift (thus,
time delay) of signals passing through the filter varies non-linearly with frequency.

66
Av

f
Figure 3.17: Frequency Response of the Butterworth characteristic
Filters with the Butterworth response are normally used when all frequencies in the passband must
have the same gain. The Butterworth response is often referred to as a maximally flat response.

3.5.2 The Chebyshev Characteristic


Filters with the Chebyshev response characteristic are useful when a rapid roll-off is required because
it provides a roll-off rate greater than – 20 dB/decade/pole. This is a greater rate than that of the
Butterworth, so filters can be implemented with the Chebyshev response with fewer poles and less
complex circuitry for a given roll-off rate. This type of filter response is characterized by overshoot
or ripples in the passband (depending on the number of poles) and an even less linear phase response
than the Butterworth.

Av

Figure 3.18: Frequency response of the Chebyshev characteristic

3.5.3 The Bessel Characteristic


Av

f
Figure 3.19: Frequency response of the Bessel characteristic
The Bessel response (also sometimes referred to as the Thomson filter) exhibits a linear phase
characteristic, meaning that the phase shift increases linearly with frequency. The result is almost no
67
overshoot on the output with a pulse input. For this reason, filters with the Bessel response are used
for filtering pulse waveforms without distorting the shape of the waveform.

In summary
For Bessel
• Flat response in the passband.
• Role-off rate less than 20dB/decade/pole.
• Phase response is linear.
• Used for filtering pulse waveforms without distorting the shape of the waveform.

For Butterworth
• Very flat amplitude, Av(dB) , response in the passband.
• Role-off rate is 20dB/decade/pole.
• Phase response is not linear.
• Used when all frequencies in the passband must have the same gain.
Often referred to as a maximally flat response

For Chebyshev
• Overshoot or ripples in the passband.
• Role-off rate greater than 20dB/decade/pole.
• Phase response is not linear - worse than Butterworth.
• Used when a rapid roll-off is required.

3.6 Review Questions


1. What determines the bandwidth of a low-pass filter?
2. How are the Q and the bandwidth of a band-pass filter related?
3. How is the selectivity affected by the Q of a filter?
4. Is the passband of a low-pass filter above or below its critical frequency?
5. How do Butterworth, Chebyshev and Bessel responses differ?
6. What are the basic parts of an active filter?
7. What determines the critical frequency of a filter?
8. What determines the roll-off rate of a filter?
9. How many poles does a second-order low-pass filter have?
10. What is the primary purpose of cascading low-pass filters?
11. How many capacitors does a two-pole filter have?
12. What roll-off rate is produced by cascading three single-pole filters?
13. Determine the critical frequency of the low-pass filter in figure 1.
0.02 𝜇𝐹
C2

+V
R2 R1
+
1.0
vin kΩ 1.0 kΩ
C1 vout
0.02 𝜇𝐹 -
Rf1
-V

1.0 kΩ Rf2

Figure 1: for Q13


68
[7.96 kHz]

14. For the 3-pole filter in figure 2, determine the capacitance values required to produce a critical
frequency of 2680 Hz if all the resistors in the RC low-pass circuits are 1.8 kΩ.

Figure 2: for Q14


[𝟎. 𝟎𝟑𝟑 𝝁𝑭]
15. A single pole high-pass filter has a frequency-selective network with 𝑅 = 2.2 𝑘Ω and 𝐶 =
0.0015 𝜇𝐹. Determine the critical frequency.

16. What is the bandwidth of a band-pass filter whose critical frequencies are 3.2 kHz and 3.9
kHz? What is the Q of this filter? Classify it as narrowband or wideband.

69
CHAPTER 4
SIGNAL GENERATION

4.1 Oscillators

Oscillators are an important class of feedback circuits that are used for signal generations, as well as
for various timing and control applications. An oscillator is a circuit that produces a periodic
waveform as output with only the dc supply voltage as a required input. This means the oscillator
continuously produces a repetitive time varying electrical signals. A repetitive input signal is not
required but is sometimes used to synchronize oscillations. The oscillator is seen as a special type of
an amplifier with infinite gain, and as a result can produce an output signal with no externally applied
input signal.
The oscillators are based on the principle of positive feedback, where a portion of the output is fed
back into the system. Positive feedback refers to the feedback in which the amplifier output signal
increases in magnitude after feedback has been applied. This feedback is cumulative. It thus also
mean that the amplified signal when fed back and passed through the amplifier becomes even
stronger.
Oscillators are designed to produce a controlled oscillation with one of two basic methods: the unity-
gain method used with sinusoidal (harmonic) oscillators and the timing method used with relaxation
oscillators.
Figure 4.1 is a block diagram that describes the oscillator. It can be described by the positive (or
regenerative) feedback system. The network used is a frequency-selective feedback, and the oscillator
is designed to produce an output even though the input is zero.

+ Amplifier gain
V0 = 0
A

+
Vi = 0

Positive
feedback Frequency-selective
feedback network
gain, 

Figure 4.1: Block diagram for a positive feedback system

As could be observed from figure 4.1, the oscillator consists of an amplifier for gain and a positive
feedback circuit that produces phase shift and attenuates the signal.
In oscillators, the electrons in the circuit are made to perform an oscillatory motion and such
electronic devices can generate AC signals of any frequency starting from a low frequency of a few
hertz to very high frequency of several gigahertz. The frequency of the generated signal depends on
the circuit constants. Thus, the output signal frequency depends on the passive components employed
in the circuit and can be varied as per the need.

70
4.2 Principle of an Electronic Oscillator

Figure 4.2: Principle of an Electronic Oscillator

By the arrangement in Figure 2(a), the capacitor, C, is charged when the switch, S, is connected to
P1. After C is fully charged, S is now connected to P2, allowing C to be discharged through the coil,
L. The current flowing through L builds an expanding magnetic field around it.
After C, has completely discharged, the magnetic field collapses. From Faraday’s laws, an Emf is
induced in L, which tries to maintain the flow of current through L in the original direction. This
electron flow charges the capacitor to opposite polarity as in Figure 2(c).
C again tries to discharge itself through L but the electron flow is now in the opposite direction and
another magnetic field in the opposite direction is built around L.

This back and forth motion of electrons in the circuit constitutes oscillations and the process
continues until all the energy given to the circuit by the battery in the course of the initial charging
of C is dissipated as heat due to 𝐼 2 𝑅 losses in the resistance of the circuit.

1
For LC circuit, 𝑓 = (4.1)
2𝜋√𝐿𝐶

It must be noted that oscillatory response was possible due to the presence of the two types of storage
elements. Having both L and C allows the flow of energy back and forth between the two.

Figure 4.3: Diminishing Amplitude

4.3 Classifications of Oscillators


All oscillators convert electrical energy from the dc power supply to periodic waveforms that can be
used for timing, control, or signal-generating applications.

They are broadly divided into two major types. These are:
(a) Sinusoidal (harmonic) (b) Non-sinusoidal (relaxation) oscillators

71
Oscillator

Figure 4.4: Output of oscillators

Both types can include active devices such as BJTs, FETs and Op-Amps and passive components
such as resistors, inductors and capacitors.
Oscillators can also be categorised on the basis of design principle used and the range of frequency
over which they are employed. According to design principle used, they could either be feedback
oscillators or negative resistance oscillators. Normally, feedback oscillators are widely employed.
Table 4.1 depicts the different types of oscillators according to operating frequencies.
Table 4.1: Types of Oscillators (Operating frequencies)
Types of Oscillators Approximate Range
Audio-frequency (AF) 20 Hz – 20 kHz
Radio-frequency (RF) 20 kHz – 30 MHz
Very low frequency (VLF) 15 – 100 kHz
Low frequency (LF) 100 – 150 kHz
Broadcast oscillators 500 kHz – 1.5 MHz
Video-frequency 0 – 5 MHz
High frequency (HF) 1.5 – 30 MHz
Very high frequency (VHF) 30 – 300 MHz
Ultra high frequency (UHF) 300 MHz – 3 GHz
Microwave oscillators Beyond 3 GHz

4.3.1 Sinusoidal or Harmonic Oscillator


The sinusoidal oscillator produces an output which has the sine waveform. In harmonic oscillators,
the energy always flows in one direction – from the active to the passive components, and the
frequency of oscillations is determined by the feedback path. It can provide output at frequencies
ranging from 20 Hz to 1 GHz. Examples of the sinusoidal are: Tuned circuits or LC feedback
oscillators, R-C oscillators, crystal oscillators etc

4.3.2 Non-Sinusoidal or Relaxation Oscillator


The non-sinusoidal produces output which rises quickly to one voltage level and later drops quickly
to another voltage level. This circuit is usually referred to as a pulse or square-wave generator. In
relaxation oscillators, the energy is exchanged between the active and passive components. Here, the
frequency is determined by time constants – specifically, the charge and discharge time constants
during the exchange of energy. Example of the waveforms are rectangular, square or sawtooth
waveform. In other words, they are of the pulse shape. Such oscillators can provide output at
frequency ranging from zero to 20 MHz.

Review Questions
1. What is an oscillator?
2. What type of feedback does an oscillator require?
72
3. What is the purpose of the feedback circuit?
4. How does a relaxation oscillator differ from a feedback oscillator?
5. Generally, which type of oscillator produces a square wave?
6. Define positive feedback.

4.4 Barkhausen Criteria


An oscillator is basically an amplifier which does not have any ac input but it operates on the principle
of positive feedback to generate an ac signal at its output. By inference, an amplifier can work as an
oscillator if positive feedback is made to exist. However, positive feedback does not always guarantee
oscillations. An amplifier will work as an oscillator if and only if it satisfies a set of conditions called
the Barkhausen Criteria. Thus, for sine wave oscillators to sustain oscillations, they must meet
Barkhausen criteria:
a) The overall loop gain of the oscillator must be unity or greater. Therefore losses can be
compensated for by an amplifying device.
b) The net phase shift (from input to output and back to the input) must be zero.
It must be noted that by deliberate design the phase shift around the loop is made 0𝑜 at the resonant
frequency. Above and below the resonant frequency, the phase shift is different from 0𝑜 . Thus
oscillations are obtained at only one frequency, the resonant frequency of the feedback circuit.

Feedback should be positive.


The voltage gain around the closed feedback loop, Acl is the product of the amplifier gain, Av and
the attenuation, β, of the feedback circuit. 𝛽, in other words is the ‘gain’ of the feedback network.
𝐴𝑐𝑙 = 𝐴𝑣 𝛽 (4.2)

If a sinusoidal wave is the desired output, a loop gain greater than 1 will rapidly cause the output to
saturate at both peaks of the waveform, producing unacceptable distortion. To avoid this, some form
of gain control must be used to keep the loop gain at exactly 1, once oscillations have started.
For example, if the attenuation of the feedback circuit is 0.01, the amplifier must have a gain of
exactly 100 to overcome this attenuation and not create unacceptable distortion (that is,
0.01 × 100 = 1.0). An amplifier gain of greater than 100 will cause the oscillator to limit both peaks
of the waveform.

4.5 Start-up Conditions


For oscillation to begin, supply of an input signal is not essential. Only the condition 𝐴𝑐𝑙 = 𝐴𝑣 𝛽 = 1
must be satisfied for self-sustained oscillations to result.
In practice, 𝐴𝑐𝑙 is made slightly greater than unity and the system starts oscillating by amplifying
noise voltage which is always present [Noise voltage is the voltage generated inside any electronic
device such as a resistor, which is due to random movement of electrons].
Thus the voltage gain around the positive feedback loop must be greater than 1 so that the amplitude
of the output can build up to a desired level. Saturation factors in the practical circuits provide an
average value of 𝐴𝑐𝑙 of 1. The resulting waveforms are never exactly sinusoidal. However, the closer
the value of 𝐴𝑐𝑙 is to exactly 1, the more nearly sinusoidal is the waveform. In other words, the gain
must decrease to 1 so that the output stays at desired level and oscillation is sustained.

When oscillation starts at to, the condition Acl >1 causes the sinusoidal output voltage amplitude to
build up to a desired level. Then Acl decreases to 1 and maintains the desired amplitude.

73
to

Acl >1 Acl = 1

Figure 4.5: Build-up of Steady State Oscillations

4.6 Sinusoidal Oscillators

4.6.1 Nature of Sinusoidal Oscillations


They are of two types, namely damped oscillations and undamped oscillations.

Damped Oscillations
They are the electrical oscillations whose amplitude decreases with time. The waveform of such type
of oscillations is shown in figure 4.6. The damped oscillations are produced by those oscillator
circuits in which power losses take place continuously during each oscillation without any means for
compensating the same. However in damped oscillations, the frequency of oscillations remains
unchanged because it depends upon the circuit parameters.

Figure 4.6: Damped Oscillations

The period for the oscillation will remain the same but the amplitude diminishes till oscillations die
down completely.

Undamped Oscillations
They are the electrical oscillations whose amplitude remains constant with time as could be observed
in figure 4.7. The undamped oscillations are produced by the oscillator circuits in which either there
are no power losses or if they have any power loss then they have provision for compensating these
losses. It must be noted that an oscillator is required to produce undamped oscillations for use in
various communication and electronics applications.

74
i

Figure 4.7: Undamped Oscillations

4.7 R-C Oscillators


These oscillators use resistors and capacitors and are used to generate low or audio-frequency
signal. Hence, they are also known as audio-frequency (AF) oscillators. They are commonly used
for generating audio-frequencies since they provide good frequency stability and waveform. Also,
with the advent of IC technology, R-C network is the only feasible solution, as it is very difficult to
make a too high value inductance in an integrated circuit. Two commonly used R-C oscillators are
R-C phase-shift oscillator and Wien-bridge oscillator.

4.7.1 Basic Working Principles of R-C Oscillators


To produce oscillations in an oscillator circuit, positive feedback is required, which means that the
voltage signal feedback should be in phase with the input signal. To provide a positive feedback at
one particular frequency, an inverting amplifier may be used with a feedback network which causes
a phase shift of 180o at the desired frequency of oscillation.

Oscillator output
at Resonant
Frequency

Amplifier with 180o Phase-shift Network


Phase Shift 𝛽∠180𝑜

Figure 4.8: Phase Shift Network

4.8 The Wien-Bridge Oscillator


The Wien-bridge oscillator is one of the most popular types of RC oscillators used in audio and sub-
audio frequency ranges. It is widely used for generating sinusoids in the frequency range up to about
1 MHz. It is an RC op-amp circuit with only a few components, easily tunable and easy to design. It
is simple in design, compact in design size and stable in its frequency output. Its output is relatively
free from distortion and its frequency can be varied easily.

4.8.1 Wien-Bridge Circuit


Figure 4.9 (a) shows the Wien-bridge which is used as the feedback network in Wien-bridge
oscillator, whilst 4.9 (b) shows the basic Wien-bridge oscillator. The amplifier in Figure 4.9 (b) can
use a transistor or FET or op-amp as amplifying device. The Wien-bridge (Figure 4.9 (a)) has four
arms. The arm AD which consists of the series combination of 𝑅1 and 𝐶1 and the arm CD which

75
consists of the parallel combination of 𝑅2 and 𝐶2 , are jointly known as the frequency sensitive arms.
This is because the components connected to the arms decide the oscillator frequency.
The resistors 𝑅3 and 𝑅4 are used to generate a reference voltage which remains constant independent
of the frequency.

Amplifier 𝑉𝑂

𝐴 Feedback
𝐴 𝐶1
𝐶1
𝑅1 𝑅3
𝑅1 𝑅3

Bridge Bridge 𝐷 𝐵
Input 𝐷 Output 𝐵
𝑅2
𝑅2 𝑅4
𝑅4 𝐶2
𝐶2
𝐶
𝐶
Feedback Network
(a) Circuit (b) Oscillator
Figure 4.9: The Wien-Bridge
The ac input voltage is applied between points A and C of the bridge. When the Wien-bridge is used
in the oscillator circuit, the feedback voltage is applied between these points as shown in Figure 4.9
(b). The ac output of the bridge is obtained between points B and D of the bridge. When used in the
oscillator circuit, these points are connected to the input of the amplifier.
Thus, the Wien-bridge circuit of Figure 4.9 (a) is used as a feedback network in the Wien bridge
oscillator circuit observed in Figure 4.9 (b).
A fundamental part of the Wien-bridge is a lead-lag circuit as shown in Figure 4.9. 𝑅1 and 𝐶1 together
form the lag portion of the circuit whilst 𝑅2 and 𝐶2 form the lead portion. The operation of this lead-
lag circuit is as follows:
At lower frequencies, the lead circuit dominates due to the high reactance of 𝐶2 . As the frequency
increases, 𝑋𝑐2 decreases, thus allowing the output voltage to increase.
At some specified frequency, the response of the lag circuit takes over, and the decreasing value of
𝑋𝑐1 causes the output voltage to decrease.
The response curve for the lead-lag circuit shown in Figure 4.10 indicates that the output voltage
peaks at a frequency called the frequency of oscillations or cut-off or resonant frequency, 𝑓𝑐 . At this
1
point the attenuation 𝑉𝑜 ⁄𝑉𝑖𝑛 (or 𝛽) of the network is if 𝑅1 = 𝑅2 and 𝑋𝑐1 = 𝑋𝑐2
3

𝑉𝑂
1
𝑉𝑂(𝑝𝑒𝑎𝑘) = 𝑉
3 𝑖𝑛

𝑓𝑂 𝑓
Figure 4.10: The Wien-Bridge Response Curve

76
1
The cut-off frequency is given by 𝑓𝑐 = (4.2)
2𝜋𝑅𝐶
1
and 𝛽 =
3

𝐶1
𝐴
𝑅1 𝑅3

𝐷 -
𝐵
𝑅2 𝑉𝑂
𝑅4 +
𝐶2
𝐶

Figure 4.11: Op-Amp Wien-Bridge Oscillator


𝑅4 𝑅3
𝐶
𝐵

𝑅3 𝑅1 -
𝐴
𝐶1 𝑉𝑂
- +
𝑅 𝐶
+ 𝑉𝑂 𝐷
𝑅4

𝑅2 𝐶2 𝐶 𝑅

Figure 4.12: Variations of Wien-Bridge Oscillator using Op-Amp

Summary of Wien Bridge Oscillator


(i) Gain of the non-inverting amplifier is adjusted at 𝐴 ≥ 3
(ii) At the cut-off frequency, the phase shift introduced by the Wien bridge is 0𝑜 and feedback
1
factor 𝛽 =
3
(iii) Wien bridge acts as the lead-lag RC feedback network. Below 𝑓𝑐 , the lead circuit
dominates and the output leads the input whiles above 𝑓𝑐 , the lag circuit dominates and
the output lags the input.
(iv) The gain of the amplifier configuration of Wien-bridge is given by:
𝑅3
𝐴 = 1+ , (4.3)
𝑅4
𝑅3 𝑅3
but from (i), 𝐴 ≥ 3 Hence 1 + ≥ 3 or ≥2
𝑅4 𝑅4
Thus 𝑅3 should be greater than two times the value of 𝑅4 , to ensure sustained oscillations.

77
Example 4.1
Determine the cut-off frequency of the Wien-bridge in figure 4.13

Figure 4.13: For example 4.1

Solution 4.1
1 1
Cut-off frequency, fc is given by 𝑓𝑐 = = = 1560 𝐻𝑧
2𝜋𝑅𝐶 2×𝜋×51×103 ×2×10−9

Example 4.2
For the Op-Amp based Wien bridge in shown in figure 4.12 (b), if the component values
are: 𝑅 = 5 𝑘Ω, 𝐶 = 5 𝑛𝐹, 𝑅3 = 15 𝑘Ω 𝑎𝑛𝑑 𝑅4 = 6 𝑘Ω,
(i) Determine whether the circuit will oscillate or not.
(ii) Obtain the resonant frequency.

4.9 The Phase Shift Oscillator


In this oscillator, oscillation occurs at the frequency where the total phase shift through the R-C
circuits is 180o. The inversion of the op-amp itself provides the additional 180o to meet the
requirement for oscillation of a 360o (or 0o) phase shift around the feedback loop.
1
The attenuation, 𝛽, of the RC feedback circuit is 𝛽 = 29.
Network analysis have shown that when necessary phase shift of 180 is obtained, this network
1
attenuates the output voltage by a factor of . This means that the amplifier must have a voltage
29
gain of 29 or more (i.e. 𝐴 ≥ 29).
1
When the amplifier voltage gain is 29 and feedback factor of R-C network, 𝛽 = 29 then the loop
1
gain is 𝐴𝛽 = × 29 = 1
29
The frequency of oscillation is given by
1
𝑓= (4.4)
2𝜋√6𝑅𝐶

4.10 Tuned Circuits or LC Feedback Oscillators


These oscillators use a tuned-circuit consisting of inductors (L) and capacitors (C) and are used to
generate high frequency signals. Hence they are also known as radio frequency (RF) oscillators. The
L-C tuned circuit causes a phase shift of 180o due to inductive or capacitive coupling in addition to
180o phase shift produced by say, a transistor. These oscillators employing L-C elements, called the
L-C oscillators are very popular for generating high frequency oscillations but they cannot be
employed for generation of low frequency oscillations because they become too bulky and expensive.
Examples of such oscillators are Armstrong or Tickler coil, Hartley, Clapp and Colpitts oscillators.
The frequency of oscillation (𝑓𝑜 ), is inversely proportional to the square root of the product of the
capacitance of the capacitor (C) and the self-inductance of the coil (L).
1
That is 𝑓𝑜 = where L is in henry and C is in farads
2𝜋√𝐿𝐶

78
4.11 Crystal Oscillators
These oscillators use quartz crystals and are used to generate highly stabilised output signal with
frequency up to 10 MHz. Thus, the crystal oscillator has a high stability in holding constant at
whatever frequency the crystal is originally cut to operate, and thus are often used whenever great
stability is required. For example, in communication transmitters and receivers, digital clocks etc.
The pierce oscillator is an example of a crystal oscillator.

4.12 RC Relaxation Oscillators


Perhaps the easiest type of oscillator to design is the RC relaxation oscillator. Its oscillatory nature is
explained by the following principle: Charge a capacitor through a resistor and then rapidly discharge
it when the capacitor voltage reaches a certain threshold voltage. After that, the cycle is repeated,
continuously. In order to control the charge/discharge cycle of the capacitor, an amplifier wired with
positive feedback is used. The amplifier acts like a charge/discharge switch. It is triggered by the
threshold voltage and also provides the oscillator with gain.

4.13 Square-Wave Oscillator


Figure 4.14 shows a simple op-amp relaxation oscillator.

Figure 4.14: Square-Wave Oscillator


When power is first applied, the op amp’s output goes toward positive saturation (it is equally likely
that the output will go to negative saturation). The capacitor will begin to charge up toward the op
amp’s positive supply voltage (around +15V) with a time constant of R1C.When the voltage across
the capacitor reaches the threshold voltage, the op amp’s output suddenly switches to negative
saturation (around −15V).The threshold voltage is the voltage set at the inverting input, which is:
𝑅3 15 𝑘Ω
𝑉𝑇 = 𝑉+ = (+15 𝑉) = +7.5 𝑉
𝑅3 +𝑅2 15 𝑘Ω+15 𝑘Ω
The threshold voltage set by the voltage divider is now −7.5 V. The capacitor begins discharging
toward negative saturation with the same R1C time constant until it reaches −7.5 V, at which time
the op amp’s output switches back to the positive saturation voltage. The cycle repeats indefinitely,
with a period equal to 2.2R1C.

4.14 Sawtooth Generator


Here’s another relaxation oscillator that generates a sawtooth waveform (see Figure 4.15). Unlike
the preceding oscillator, this circuit resembles an op amp integrator network; with the exception of
the PUT (programmable unijuction transistor) in the feedback loop. The PUT is the key ingredient
that makes this circuit oscillates. Here’s how this circuit works.

79
Figure 4.15: Sawtooth Generator

Assuming the circuit does not contain the PUT. In this case, the circuit would resemble a simple
integrator circuit; when a negative voltage is placed at the inverting input (−), the capacitor charges
up at a linear rate toward the positive saturation voltage (+15 V). The output signal would simply
provide a one-shot ramp voltage, that is, it would not generate a repetitive triangular wave. In order
to generate a repetitive waveform, we must now include the PUT. The PUT introduces oscillation
into the circuit by acting as an active switch that turns on (anode-to-cathode conduction) when the
anode-to-cathode voltage is greater than its gate voltage. The PUT will remain on until the current
through it falls below the minimum holding current rating. This switching action acts to rapidly
discharge the capacitor before the output saturates. When the capacitor discharges, the PUT turns off,
and the cycle repeats. The gate voltage of the PUT is set via voltage-divider resistors R4 and R5.The
R1 and R2 voltage-divider resistors set the reference voltage at the inverting input, while the diodes
help stabilize the voltage across R2 when it is adjusted to vary the frequency. The output voltage
amplitude is determined by R4, while the output frequency is approximated by the expression below
the figure. (The 0.5 V represents a typical voltage drop across a PUT.)

4.15 Triangular –Wave/ Square – Wave Generator


Figure 4.16 is a simple dual op amp circuit that generates both triangular and square waveforms. This
circuit combines a triangle-wave generator with a comparator. The rightmost op-amp in the circuit
acts as a comparator—it is wired with positive feedback. If there is a slight difference in voltage
between the inputs of this op amp, the V2 output voltage will saturate in either the positive or negative
direction.

Figure 4.16: Triangular Wave / Square Wave Generator


For example, if the op amp saturates in the positive direction, it will remain in that saturated state until
the voltage at the non-inverting input (+) drops below the negative threshold voltage (−VT), at which
time V2 will be driven to negative saturation. The threshold voltage is given by:
𝑅3
𝑉𝑇 = 𝑉 (4.5)
𝑅3 +𝑅2 𝑠𝑎𝑡

80
where Vsat is a volt lower than the op amp’s supply voltage. Now this comparator is used with a ramp
generator (leftmost op amp section).The output of the ramp generator is connected to the input of the
comparator, while its output is fed back to the input of the ramp generator. Each time the ramp voltage
reaches the threshold voltage, the comparator changes states. This gives rise to oscillation. The period
of the output waveform is determined by the R1C time constant, the saturation voltage, and the
threshold voltage:
4𝑉𝑇
𝑇=𝑉 × 𝑅1 𝐶 (4.6)
𝑠𝑎𝑡

Where frequency is 1/T.

4.16 Unijunction Oscillator


Op-amps are not the only active ingredient used to construct relaxation oscillators. Other components,
such as transistors and digital logic gates, can take their place. Here is a unijunction transistor (UJT),
along with some resistors and a capacitor, which makes up a relaxation oscillator that is capable of
generating three different output waveforms.

Figure 4.17: Unijunction Oscillator

During operation, at one instant in time, C charges through R until the voltage present on the emitter
reaches the UJT’s triggering voltage (refer to figure 4.17). Once the triggering voltage is exceeded,
the E-to-B1 conductivity increases sharply, which allows current to pass from the capacitor-emitter
region through the emitter-base 1 region and then to ground. When this occurs, C suddenly loses its
charge, and the emitter voltage suddenly falls below the triggering voltage. After that, the cycle
repeats itself. The resulting waveforms generated during this process are shown in the figure. The
frequency of oscillation is given by:
1
𝑓= (4.7)
𝑅𝑔 𝐶𝑔 ln[1⁄(1−𝜂)]

where η is the UJT’s intrinsic standoff ratio, which is typically around 0.5.

4.17 Digital Oscillator


A simple relaxation oscillator can be built from a Schmitt trigger inverter IC and an RC network.
Schmitt triggers are used to transform slowly changing input waveforms into sharply defined jitter-
free output waveforms.
When power is first applied to the circuit, the voltage across C is zero, and the output of the inverter
is high (+5 V). The capacitor starts charging up toward the output voltage via R. When the capacitor
voltage reaches the positive-going threshold of the inverter (e.g., 1.7 V), the output of the inverter
goes low (towards 0 V). With the output low, C discharges toward 0 V.
When the capacitor voltage drops below the negative-going threshold voltage of the inverter (e.g.,
0.9 V), the output of the inverter goes high. The cycle repeats.
The on/off times are determined by the positive- and negative-going threshold voltages and the RC
time constant.
81
(a) Using Schmitt trigger inverter (b) Using inverters
Figure 4.18: Digital Oscillator
The third example is a pair of CMOS inverters that are used to construct a simple square wave RC
relaxation oscillator. The circuit can work with voltages ranging from 4 to 18 V. The frequency of
oscillation is given by:

(4.8)
R can be adjusted to vary the frequency. All relaxation oscillators shown in this section are relatively
simple to construct. Now, as it turns out, there is even an easier way to generate basic waveforms.
The easy way is to use an IC especially designed for the task. An incredibly popular square wave-
generating chip that can be programmed with resistors and a capacitor is the 555 timer IC.

4.18 The 555 Timer IC


The 555 timer IC is a very useful precision timer that can act as either a timer or an oscillator. In timer
mode, better known as monostable mode, the 555 simply acts as a “one-shot” timer; when a trigger
voltage is applied to its trigger lead, the chip’s output goes from low to high for a duration set by an
external RC circuit. In oscillator mode, better known as astable mode, the 555 acts as a rectangular-
wave generator whose output waveform (low duration, high duration, frequency, etc.) can be adjusted
by means of two external RC charge/discharge circuits.
The 555 timer IC is easy to use (requires few components and calculations) and inexpensive and can
be used in an amazing number of applications. For example, with the aid of a 555, it is possible to
create digital clock waveform generators, LED and lamp flasher circuits, tone-generator circuits
(sirens, metronomes, etc.), one-shot timer circuits, bounce-free switches, triangular-waveform
generators, frequency dividers, etc.

4.18.1 How a 555 Works (Astable Operation)


Figure 4.19 is a simplified block diagram showing what is inside a typical 555 timer IC. The overall
circuit configuration shown here (with external components included) represents the astable 555
configuration.
The 555 gets its name from the three 5-kΩ resistors shown in the block diagram. These resistors act
as a three-step voltage divider between the supply voltage (VCC) and ground. The top of the lower 5-
kΩ resistor (+ input to comparator 2) is set to 1⁄3VCC, while the top of the middle 5-kΩ resistor (−
input to comparator 2) is set to 2⁄3VCC. The two comparators output either a high or low voltage based
on the analog voltages being compared at their inputs.
If one of the comparator’s positive inputs is more positive than its negative input, its output logic
level goes high; if the positive input voltage is less than the negative input voltage, the output logic
level goes low. The outputs of the comparators are sent to the inputs of an SR flip-flop. The flip-flop
looks at the R and S inputs and produces either a high or a low based on the voltage states at the
inputs.
82
Figure 4.19: A 555 Timer

4.18.2 Pins of the 555 Timer


Pin 1 (ground). IC ground
Pin 2 (trigger). Input to comparator 2, which is used to set the flip-flop. When the voltage at pin 2
crosses from above to below 1⁄3VCC, the comparator switches to high, setting the flip-flop.
Pin 3 (output). The output of the 555 is driven by an inverting buffer capable of sinking or sourcing
around 200 mA. The output voltage levels depend on the output current but are approximately
Vout,(high) = VCC − 1.5 V and Vout,(low) = 0.1 V.
Pin 4 (reset). Active-low reset, which forces Q high and pin 3 (output) low.
Pin 5 (control). Used to override the 2⁄3VCC level, if needed, but is usually grounded via a 0.01-μF
bypass capacitor (the capacitor helps eliminate VCC supply noise). An external voltage applied here
will set a new trigger voltage level.
Pin 6 (threshold). Input to the upper comparator, which is used to reset the flip-flop. When the voltage
at pin 6 crosses from below to above 2⁄3VCC, the comparator switches to a high, resetting the flip-
flop.
Pin 7 (discharge). Connected to the open collector of the npn transistor. It is used to short pin 7 to
ground when Q is high (pin 3 low). This causes the capacitor to discharge.
Pin 8 (Supply voltage VCC). Typically between 4.5 and 16 V for general-purpose TTL 555 timers.
(For CMOS versions, the supply voltage may be as low as 1 V.)

In the astable configuration, when power is first applied to the system, the capacitor is uncharged.
This means that 0 V is placed on pin 2, forcing comparator 2 high. This in turn sets the flip-flop so
that Q is high and the 555’s output is low (a result of the inverting buffer). With Q high, the
discharge transistor is turned on, which allows the capacitor to charge toward VCC through R1 and R2.
When the capacitor voltage exceeds 1⁄3VCC, comparator 2 goes low, which has no effect on the SR
flip-flop. However, when the capacitor voltage exceeds 2⁄3VCC, comparator 1 goes high, resetting the
flipflop and forcing Q high and the output low. At this point, the discharge transistor turns on and
shorts pin 7 to ground, discharging the capacitor through R2. When the capacitor’s voltage drops
below 1⁄3VCC, comparator 2’s output jumps back to a high level, setting the flip-flop and making Q
low and the output high. With Q low, the transistor turns on, allowing the capacitor to start charging
again. The cycle repeats over and over again. The net result is a square wave output pattern whose

83
voltage level is approximately VCC − 1.5 V and whose on/off periods are determined by the C, R1 and
R2.

4.18.3 Basic Astable Operation


When a 555 is set up in astable mode, it has no stable states; the output jumps back and forth. The
time duration Vout remains low (around 0.1V) is set by the R1C time constant and the 1⁄3VCC and
2⁄3VCC levels; the time duration Vout stays high (around VCC − 1.5 V) is determined by the (R1 + R2)
C time constant and the two voltage levels (see graphs in Figure 4.20).
After doing some basic calculations, the following two practical expressions arise:
𝑡𝑙𝑜𝑤 = 0.693𝑅2 𝐶 (4.9)
𝑡ℎ𝑖𝑔ℎ = 0.693(𝑅1 + 𝑅2 )𝐶 (4.10)
The duty cycle (the fraction of the time the output is high) is given by:
𝑡ℎ𝑖𝑔ℎ
Duty cycle = (4.11)
𝑡ℎ𝑖𝑔ℎ +𝑡𝑙𝑜𝑤

And the frequency of the output waveform is:


1
𝑓= (4.12)
𝑡ℎ𝑖𝑔ℎ +𝑡𝑙𝑜𝑤

For example, using the circuit in Figure 4.20 (a)

(a) Circuit (b) Equivalent Graphs


Figure 4.20: Basic Astable Operation

𝑡𝑙𝑜𝑤 = 0.693(2 × 104 )(680 × 10−9 ) = 9.6 𝑚𝑠


𝑡ℎ𝑖𝑔ℎ = 0.693(1 × 104 + 2 × 104 )(680 × 10−9 ) = 14.1 𝑚𝑠
1
𝑓= = 42 𝐻𝑧
14.1+9.6
14.1
Duty cycle = = 0.6
14.1+9.6

For reliable operation, the resistors should be between approximately 10 kΩ and 14 MΩ, and the
timing capacitor should be from around 100 pF to 1000 μF. The graph will give you a general idea
of how the frequency responds to the component values.

84
4.18.4 Low-Duty-Cycle Operation (Astable Mode)
Now there is a slight problem with the last circuit, i.e. you cannot get a duty cycle that is below 0.5
(or 50 percent). In other words, you cannot make thigh shorter than tlow. For this to occur, the R1C
network (used to generate tlow) would have to be larger the (R1 + R2)C network (used to generate thigh).
Simple arithmetic tells us that this is impossible; (R1 + R2)C is always greater than R1C.

Figure 4.21: Circuit to Depict Low-Duty Cycle Operation


Astable Operation
How is this situation remedied? You attach a diode across R2, as shown in figure 4.21. With the diode
in place, as the capacitor is charging (generating thigh), the preceding time constant (R1 + R2) C is
reduced to R1C because the charging current is diverted around R2 through the diode. With the diode
in place, the high and low times become
𝑡ℎ𝑖𝑔ℎ = 0.693(1 × 104 )(1 × 10−6 ) = 6.9 𝑚𝑠 (from 𝑡ℎ𝑖𝑔ℎ = 0.693𝑅1 𝐶)
𝑡𝑙𝑜𝑤 = 0.693(47 × 103 )(1 × 10−6 ) = 32.5 𝑚𝑠 (from 𝑡𝑙𝑜𝑤 = 0.693𝑅2 𝐶)
1 6.9
𝑓= = 25 𝐻𝑧 and Duty cycle = = 0.18
6.9 𝑚𝑠+9.6 𝑚𝑠 6.9+32.5
To generate a duty cycle of less than 0.5 or 50%, simply make 𝑅1 less than 𝑅2

4.18.5 How a 555 Works (Monostable Operation)

Figure 4.22 A 555 Timer in Monostable Configuration (One-shot Mode)

85
Figure 4.22 shows a 555 hooked up in the monostable configuration (one-shot mode). Unlike the
astable mode, the monostable mode has only one stable state. This means that for the output to switch
states, an externally applied signal is needed.
In the monostable configuration, initially (before a trigger pulse is applied) the 555’s output is low,
while the discharge transistor in on, shorting pin 7 to ground and keeping C discharged. Also, pin 2
is normally held high by the 10-k pull-up resistor.
Now, when a negative-going trigger pulse (less than 1⁄3VCC) is applied to pin 2, comparator 2 is
forced high, which sets the flip-flop’s Q to low, making the output high (due to the inverting buffer),
while turning off the discharge transistor.

Figure 4.23 Basic Monostable Operation

This allows C to charge up via R1 from 0 V toward VCC. However, when the voltage across the
capacitor reaches 2⁄3VCC, comparator 1’s output goes high, resetting the flip-flop and making the
output low, while turning on the discharge transistor, allowing the capacitor to quickly discharge
toward 0 V. The output will be held in this stable state (low) until another trigger is applied.
The monostable circuit only has one stable state. That is, the output rests at 0 V (in reality, more like
0.1 V) until a negative-going trigger pulse is applied to the trigger lead—pin 2. (The negative-going
pulse can be implemented by momentarily grounding pin 2, say, by using a pushbutton switch
attached from pin 2 to ground.) After the trigger pulse is applied, the output will go high (around VCC
− 1.5 V) for the duration set by the R1C network.Without going through the derivations, the width of
the high output pulse is twidth = 1.10R1C For reliable operation, the timing resistor R1 should be
between around 10 kΩ and 14 MΩ, and the timing capacitor should be from around 100 pF to 1000
μF.

4.19 Voltage-Controlled Oscillators

Besides the 555 timer IC, there are a number of other voltage-controlled oscillators (VCOs) on the
market—some of which provide more than just a square wave output. For example, the NE566
function generator is a very stable, easy-to-use triangular wave and square wave generator. In the 556
circuit below, R1 and C1 set the centre frequency, while a control voltage at pin 5 varies the frequency;
the control voltage is applied by means of a voltage-divider network (R2, R3, R4). The output
frequency of the 556 can be determined by using the formula shown in Figure 4.24.

86
2(𝑉𝑐𝑐 −𝑉𝑖𝑛 )
𝑓=
𝑅1 𝐶1 𝑉𝑐𝑐
0.75𝑉𝑐𝑐 ≤ 𝑉𝑐 ≤ 𝑉𝑐𝑐
2𝐾 < 𝑅1 < 20 𝐾
𝑉𝑐 is set by the voltage divider 𝑅1 , 𝑅2 and 𝑅3

Other VCOs, such as the 8038 and the XR2206, can create a trio of output waveforms, including a
sine wave (approximation of one, at any rate), a square wave, and triangular wave.

Figure 4.24: The 556 Timer

Some VCOs are designed specifically for digital waveform generation and may use an external crystal
in place of a capacitor for improved stability.

87

You might also like