Lecture Material 2023
Lecture Material 2023
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+
V- 4 5 Balance
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.
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
V1 i1 -
V2 i2 +
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 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.
4
i1
V1 -
Ri AVin Vo
Vin n
Ro
V2
i2 +
-
Vi Vo
+
(a) Non-inverting
𝑅
𝑉2
+
𝑉𝑂
𝑉1 -
𝑣𝑖
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
𝑅
𝑉1
-
𝑉𝑂
+
𝑉2
𝑣𝑖
From equation (1.9), the output is A times larger than the input and out of phase with the input.
6
𝑅1
𝑉1
-
𝑉𝑂
+
𝑉2
𝑣𝑖1 𝑅2
𝑅𝐿
𝑣𝑖2
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
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
+
Vi Vo
-
Negative
feedback
network
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.
8
Review Questions
a) What are the connections to a basic op-amp?
c) How does the voltage gain of a practical op-amp differ from an ideal op-amp?
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
𝐼𝑓 𝑅𝑓
𝐼𝑖 𝑅𝑖
𝑷 𝑖− 𝑉1
-
𝑉𝑔 𝑉𝑂
𝑖+
𝑉𝑖 +
𝑉2
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.
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
Example 1.2
𝑅𝑓
5 𝑘Ω
- 𝑉𝑂
+
𝑉𝑖
(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
𝑅𝑓
+
𝑉𝑂
- 𝑅𝑖
𝑉𝑖 𝑅𝑓 -
𝑉𝑂
+
𝑅𝑖 𝑉𝑖
-
𝑉𝑂
+
𝑉𝑖
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
Solution 1.3
𝑅𝑖 = 5 𝑘Ω; 𝑅𝑓 = 60 𝑘Ω; 𝑉𝑖 = 40 𝑚𝑉
𝑅𝑓 60
(i) 𝐴 = 1+ = 1+ = 13
𝑅𝑖 5
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
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
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
+Vo(max)
Vo t
-Vo(max)
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
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)
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
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
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
+ +
V1
Stage 1 Stage 2 V3 = A2V2 Stage 3
V2 = A1V1 Vo = A3V3
- A1 A2 A3 -
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
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
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
𝑅𝑖
-
𝑉𝑂
+
𝑉1
𝑉2
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
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
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
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
Solution 1.9
For the difference amplifier,
𝑅 𝑅 𝑅2
𝑉𝑜 = (𝑅2 + 1) (𝑅 +𝑅
4
) 𝑉2 − 𝑉
1 3 4 𝑅1 1
21
𝐶
𝐼𝐶
𝑅
-
𝐼𝑅 𝑉𝑂
+
𝑉𝑖
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
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
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
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
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]
P.2. Calculate the input voltage, V1, if 𝑅1 = 100 Ω; 𝑅𝑓 = 1 𝑘Ω, and 𝑉𝑜𝑢𝑡 = 550 𝑚𝑉
25
P.5 Calculate the input voltage if the final output, Vo, is 20.40 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
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
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.
+ + 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.
Output
Transducer
or Sensor
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.
Disadvantages
a. Closed loop systems are expensive
b. Maintenance difficult
c. Complicated installation
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
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 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
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
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.
Solution 2.3
At frequency, 𝑓1 , the gain A is 5000180 and the feedback factor is 0.020. 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
For f2,
4500160 4500160 4500160
Af = = =
1 − 81155 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,
100065 100065 100065
Af = = =
1 − 1.14855 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,
50020 50020 50020
Af = = =
1 − 0.55 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.
Error
Reference + signal System gain Output
Input, R(s) E(s)
G(s)
C(s)
--
B(s)
Sensing system
gain
H(s)
35
𝐸(𝑠) = 𝑅(𝑠) − 𝐵(𝑠) (2.12)
𝐸(𝑠) 1
= (2.15)
𝑅(𝑠) 1+𝐺(𝑠)𝐻(𝑠)
𝐵(𝑠) 𝐺(𝑠)𝐻(𝑠)
Primary feedback ratio = = (2.16)
𝑅(𝑠) 1+𝐺(𝑠)𝐻(𝑠)
R(s) + C (s)
G(s)
-
B (s)
𝐶(𝑠) 𝐺(𝑠)
=
𝑅(𝑠) 1+𝐺(𝑠)
36
2.6 Cascaded Transfer Functions
𝐶1 (𝑠) 𝐶(𝑠)
= 𝐺1 (𝑠) and = 𝐺2 (𝑠)
𝑅(𝑠) 𝐶1 (𝑠)
R(s) C(s)
G1(s).G2(s)
G1(s)
+
+
R(s) G2(s) C(s)
+
G3(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.
≡
G(s)
Figure 2.8: Take-off point ahead of block
𝑅2 (𝑆)
𝐶(𝑠) = 𝑅1 (𝑆)𝐺(𝑆) ± 𝑅2 (𝑆) 𝐶(𝑠) = [𝑅1 (𝑆) ± ] 𝐺(𝑆)
𝐺(𝑆)
= 𝑅1 (𝑆)𝐺(𝑆) ± 𝑅2 (𝑆)
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)
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
Solution 2.6
-
R(s) + + G1 + G2 G3 C(s)
- -
𝐻1
𝐺3
a: For step 1
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
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.
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.
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).
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.
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
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 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.
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.
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
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
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).
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.
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
S6 1 3 5 5
S5 2 4 8
S4 1 1 5
S3 2 -2
S2 2 5
S1 -7
S0 5
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.
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]
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.
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
3.3.2 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
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
.
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
+V
R1
+
vin
C1 vout
- Rf1
-V
Rf2
58
3.3.9 Cascaded System
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.
❖ Phase shift refers to the phase difference between a high and low pass filter set for the same
roll-off frequency.
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
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
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
Stage 1 Stage 2
C2
+V
R2 R1
+ +V
vin R3
+
C1
- Rf1 C3 vout
-V - Rf3
-V
Rf2
Rf4
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
R2
+V
C2 C1
+
vin
R1 vout
- Rf1
-V
Rf2
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
C2
Rf
+V
R1 C1
-
vin R2
vout
+
-V
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.
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.
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.
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.
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.
+V
R2 R1
+
1.0
vin kΩ 1.0 kΩ
C1 vout
0.02 𝜇𝐹 -
Rf1
-V
1.0 kΩ Rf2
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.
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,
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
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.
They are broadly divided into two major types. These are:
(a) Sinusoidal (harmonic) (b) Non-sinusoidal (relaxation) oscillators
71
Oscillator
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
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.
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.
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
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.
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
Oscillator output
at Resonant
Frequency
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
𝑅3 𝑅1 -
𝐴
𝐶1 𝑉𝑂
- +
𝑅 𝐶
+ 𝑉𝑂 𝐷
𝑅4
𝑅2 𝐶2 𝐶 𝑅
77
Example 4.1
Determine the cut-off frequency of the Wien-bridge in figure 4.13
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.
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.
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.)
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
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.
(4.8)
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.
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.
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.
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.
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.
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.
Some VCOs are designed specifically for digital waveform generation and may use an external crystal
in place of a capacitor for improved stability.
87