Вы находитесь на странице: 1из 95

ELEC2134 Notes

Circuits & Signals

By Jason Ha
(B.Eng. Electrical Engineering, UNSW)
Contents

1 Background, Ohm’s Law, Methods of Circuit Analysis 5

2 Circuit Theorems 6

3 Capacitors, Inductors and Operational Amplifiers 7


3.1 Capacitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.2 Capacitor In A Circuit Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.3 Current Through A Capacitor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.4 Inductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.5 Inductors In A Circuit Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.6 Operational Amplifiers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.6.1 Ideal Op-Amp Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.7 Inverting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.7.1 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.8 Non-inverting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.8.1 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.9 Voltage follower . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.10 Summing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.10.1 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.11 Difference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.11.1 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.12 Cascading Op-Amps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.12.1 Cascade Example 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.13 Differentiator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.13.1 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.14 Integrator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.15 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

4 Transient Response In First Order RLC Circuits 22


4.1 Transient Behaviour During Discharge In An RC Circuit . . . . . . . . . . . . . . . . . . . 22
4.2 Transient Behaviour During Charge Phase In An RC Circuit Due To Step-Response From
Source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.2.1 The Unit-Step Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.2.2 Role Of Unit-Step In RC Step-Response . . . . . . . . . . . . . . . . . . . . . . . . 25
4.3 Short-Cut Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.4 Finding vc (∞) and vc (0) via Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.5 Summary Of Steps To Determine Capacitor Voltage From Step-Response . . . . . . . . . 29
4.6 Transient Behaviour During Discharge In An RL Circuit . . . . . . . . . . . . . . . . . . . 30
4.6.1 Voltage across the inductor in a source-free RL circuit . . . . . . . . . . . . . . . . 31
4.6.2 Voltage across the resistor in a source-free RL circuit . . . . . . . . . . . . . . . . . 31

2
4.7 Transient Behaviour During Charge Phase In An RL Circuit Due To Source Excitation . 32
4.8 Finding i(0) and i(∞) via Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.9 Summary Of Steps To Determine Circuit (Inductor) Current From Step-Response . . . . 36
4.10 RL-Analysis Using Calculus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

5 Transient Response In Second Order RLC Circuits 39


5.1 Source-Free RLC Series Circuit: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.1.1 Overdamped Case (α > ω0 ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
5.1.2 Critically-Damped (α = ω0 ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
5.1.3 Underdamped Response (α < ω0 ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
5.2 Source-Free RLC Parallel Circuit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.3 Step-Response In An RLC Circuit + Voltage Source . . . . . . . . . . . . . . . . . . . . . 45
5.3.1 Step-Response In Series RLC Circuit . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.3.2 Step-Response In A Parallel RLC Circuit + Current Source . . . . . . . . . . . . . 46

6 Sinusoidal Signals, Sinusoidal Steady-State Analysis Of Circuits 47

7 AC Power Analysis 48
7.1 Instantaneous Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
7.2 Average Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
7.3 Average Power (Frequency Domain) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
7.3.1 Special Cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
7.4 Maximum Average Power Transfer In AC . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
7.4.1 Example Of Maximum Average Power Transfer Analysis . . . . . . . . . . . . . . . 53
7.5 Effective/RMS Value . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
7.5.1 RMS-Amplitude Relationship . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
7.6 Example RMS Question . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
7.7 Apparent Power & Power Factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
7.7.1 What Leading and Lagging Mean . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
7.8 Example Question Of Apparent Power & Power Factor . . . . . . . . . . . . . . . . . . . . 59
7.9 Complex Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
7.9.1 Average VS Reactive Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
7.9.2 Leading VS Lagging From The Sign Of Reactive Power . . . . . . . . . . . . . . . 60
7.10 Conservation Of AC Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
7.10.1 Conservation Of AC Power In Parallel Configuration . . . . . . . . . . . . . . . . . 61
7.10.2 Conservation Of AC Power In Series Configuration . . . . . . . . . . . . . . . . . . 61
7.11 Power Factor Correction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

8 Magnetically Coupled Circuits 64


8.1 Mutual Inductance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
8.1.1 Mutual Inductance Relationship . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
8.2 Mutual Inductance: Sign Convention . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
8.3 Mutual Inductance: Circuit Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
8.4 Mutual Inductance: General Circuit Analysis . . . . . . . . . . . . . . . . . . . . . . . . . 68
8.5 Mutual Inductance: Sum Of Inductances . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
8.5.1 Series-Aiding Connection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
8.5.2 Series-Opposing Connection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
8.6 Energy In A Coupled Circuit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
8.7 Coupling Coefficient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
8.7.1 Example Coupling Coefficient Question . . . . . . . . . . . . . . . . . . . . . . . . 72
8.8 Transformers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
8.9 Input Impedance, Zin & Reflected Impedance, ZR . . . . . . . . . . . . . . . . . . . . . . 74
8.10 Ideal Transformer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
8.11 Eliminating Transformer By Reflection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
8.12 Step-Up/Step-Down Transformers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
8.13 Autotransformer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

9 Frequency Response, Transfer Function, Resonant Circuits 78


9.1 Frequency Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
9.2 Transfer Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
9.3 Examples Of Transfer Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
9.4 Transfer Function: Example Question . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
9.5 Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
9.6 Series RLC Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
9.6.1 Obtaining The Resonant Frequency, ωo . . . . . . . . . . . . . . . . . . . . . . . . 81
9.6.2 Maximum & Half-Power Frequencies . . . . . . . . . . . . . . . . . . . . . . . . . . 82
9.6.3 Bandwidth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
9.6.4 Quality Factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
9.7 Parallel RLC Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
9.8 Half-Power Frequencies Independent Of Circuit Configuration . . . . . . . . . . . . . . . . 84
9.9 Summary Of Important Equations For Resonant RLC Circuits . . . . . . . . . . . . . . . 84
9.10 Filters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
9.11 Passive Filters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
9.12 Low-Pass Filter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
9.13 High-Pass Filter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
9.14 Bandpass Filter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
9.15 Bandstop Filter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
9.16 Identifying The Type Of Filter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
9.17 General Case For Transfer Function At Corner/Centre Frequency . . . . . . . . . . . . . . 91
9.18 Active Filters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

10 Introduction To Laplace Transform, Applications And State Variables 93

11 Fourier Series & Fourier Transform 94

12 Two-Port Networks 95
Chapter 1

Background, Ohm’s Law, Methods of


Circuit Analysis

5
Chapter 2

Circuit Theorems

6
Chapter 3

Capacitors, Inductors and Operational


Amplifiers

3.1 Capacitors
The capacitor is a circuit element that stores charge over time. More often, we say that the capacitor
stored voltage. The capacitor consists of two parallel electrical plates of equal area with one plate con-
nected to the positive terminal and the other to the negative terminal.

The amount of charge stored is proportional to the voltage stored in the capacitor via a property called
capacitance, C:

Q = CV (3.1)
The capacitance of the capacitor is given by:

A
C= (3.2)
d
Where  is the dielectric constant associated with the material between the plates. If its simply air, then
 = 1.000.

7
3.2 Capacitor In A Circuit Diagram
The capacitor is denoted by ”| |” in a circuit diagram. The polarity of each end is dependent on which
terminal of the source it is connected to. An example of one in a circuit diagram is:

If the capacitor is connected to the circuit in such a fashion, its stored voltage is not constant, and will
vary with time. i.e.

q(t) = Cvc (t) (3.3)

3.3 Current Through A Capacitor


As we know, current is defined as the rate at which charge moves per unit time. i.e we express this as:

dq(t)
i(t) = (3.4)
dt

Since we know the charge stored in the capacitor, we can differentiate equation (3.3) to obtain the current
as a function of time:

 
dq(t) d dvc (t)
i(t) = = Cvc (t) = C (C is constant) (3.5)
dt dt dt

3.4 Inductors
The inductor, in the form of a conductive coil, works similarly to the capacitor but stores electrical energy
in the magnetic field it produces. The inductance is directly proportional to the magnetic field, φ and
inversely proportional to the current, i:

φ
L= (3.6)
i
The voltage drop across an inductor can be determined as the rate of change of the magnetic field
experienced by the inductor:

dφ d(Li) di
vL (t) = = =L (3.7)
dt dt dt
3.5 Inductors In A Circuit Diagram
Below is a diagram of the circuit notation for an inductor:

As stated on the previous page, the voltage drop across an inductor is given by:

di(t)
vL (t) = L (3.8)
dt

The current through the inductor can be found through integration. Rearranging (3.8) and integrating,
we find that:
Z t
1 t
Z
di(t) = vL (t)dt (3.9)
0 L 0

1 t
Z
i(t) − i(0) = vL (t)dt (3.10)
L 0
Hence, the current through an inductor at any time, t, is given by:

Z t
1
i(t) = i(0) + vL (t)dt (3.11)
L 0
3.6 Operational Amplifiers
Operational amplifier’s (or op-amps) are active circuit elements that act as dependent (controlled) voltage
sources. They perform mathematical operations such as:
• Addition & subtraction
• Multiplication & division
• Integration & differentiation
• Summing & difference
The circuit symbol for an ideal op-amp is shown below, characterised for its inverting (-ve) and non-
inverting (+ve) terminals:

Figure 3.1: Circuit symbol for the op-amp

3.6.1 Ideal Op-Amp Characteristics


The ideal op-amp should exhibit the following properties:
• Infinite input resistance (Rin → ∞)
• Zero output resistance (Rout → 0)
• Infinite Open-Loop Voltage Gain (A ≈ ∞)
These ideal characteristics are achieve by the following criteria:
• i− = i+ = 0 (i.e. no current enters the terminals)
• v+ = v− (i.e. the voltage at each terminal MUST be equal)
With these characteristics, the different types of op-amps that are used in circuits are:
• Inverting
• Non-inverting
• Voltage follower
• Summing
• Difference
• Integrator
• Differentiator
3.7 Inverting
The inverting op-amp essentially takes an input voltage that is, say, increasing (+ve) and produces an
output voltage that is negatively proportional to the input voltage (-ve, decreasing):

Figure 3.2: Inverting Op-Amp

i.e. the input-output voltage relationship is given by:

 
Rf
vout = Avin = − vin (3.12)
Rin

3.7.1 Derivation
At node 2 (v2 ), the non-inverting (+) terminal is directly connected to ground. Which means v+ = 0.
Using the ideal op-amp criteria:

v+ = v− = 0 (3.13)
Now, we perform KCL at node 1 (v1 ). Noting the direction of the currents in/out of the node:
v1 − v− v− − vo
= (3.14)
Ri Rf
v1 −vo
= (3.15)
Ri Rf
Rearranging and solving for vo , we get the relation:

Rf
vo = − vi (3.16)
Ri
3.8 Non-inverting
The non-inverting op-amp, contrary to the inverting variant, produces an output voltage that is directly
proportional to the input voltage (i.e. +ve in, +ve out).

Figure 3.3: Non-Inverting Op Amp

The open-loop gain gives rise to the input-output voltage relationship (NOTE: The gain does not have
the same form as the inverting scenario):

 
Rf
vout = Avin = 1+ vin (3.17)
Ri

3.8.1 Derivation
Observe the non-inverting terminal (node 2, v2 ). Due to the voltage source, vi , directly connected to the
terminal, v+ = vi . Using the ideal op-amp criteria:

v− = v+ = vi (3.18)
Performing KCL at node 1 (v1 ) and noting the current going in/out of the node:
0 − v− v− − vo
= (3.19)
Rin Rf
−vi vi − vo
= (3.20)
Rin Rf
Rearranging and solving for vo , we get that relation:
 
Rf
vo = 1 + vin (3.21)
Ri
3.9 Voltage follower
The voltage follower op-amp is is used to produce an output voltage that is equal to the input voltage.
This is used to isolate one portion of a circuit from another portion (esp. in a cascaded op-amp configu-
ration) due to its high-impedance. This prevents any interstage loading that can occur in the cascaded
configuration.

Figure 3.4: Voltage Follower (Variation Of The Inverting Op-Amp)

The input-output voltage relationship is the following:

vo = vi (3.22)

Its use as a separator in a circuit with two portions (stages) is shown below:

Figure 3.5: Separating stages in a cascaded circuit


3.10 Summing
The summing op-amp is an inverting op-amp but with multiple voltage inputs. This is apparent from
the input-output voltage equation where the gain is in the same form as that from the inverting op-amp:

Figure 3.6: Summing Op-Amp

The input-output voltage relationship is:

 
Rf Rf Rf
vo = − v1 + v2 + v3 vi (3.23)
R1 R2 R3

3.10.1 Derivation
Again, observe the non-inverting terminal directly connected to the ground. Hence, v+ = 0. Using the
ideal op-amp criteria:

v+ = v− = 0 (3.24)
Now, performing KCL at node a (v− ), we node that we have three currents entering the node from the
input voltages and one leaving the node through Rf . Hence:

i1 + i2 + i3 = i (3.25)
In terms of voltages:
v1 − v− v2 − v− v3 − v− v− − vo
+ + = (3.26)
R1 R2 R3 Rf
v1 v2 v3 −vo
+ + = (3.27)
R1 R2 R3 Rf
Rearranging and solving the vo , we get the expression we expected:
 
Rf Rf Rf
vo = − v1 + v2 + v3 vi (3.28)
R1 R2 R3
3.11 Difference
The difference op-amp is used to amplify the difference between two input voltage (signals) but eliminates
any signals common between the two inputs.

Figure 3.7: Difference Op-Amp

The input-output voltage relationship is given below:

    
R2 1 + R1 /R2 R2
vo = v2 − v1 (3.29)
R1 1 + R3 /R4 R1

3.11.1 Derivation
Observe node b (vb = v+ ). Performing KCL, we get:
v2 − v+ vb − 0
= (3.30)
R3 R4
 
R3
v+ = v2 (3.31)
R3 + R4
Using the ideal op-amp relation, we know that:
 
R3
v+ = v− = v2 (3.32)
R3 + R4
This is useful when we perform KCL at node a (va ):
v1 − v− v− − vo
= (3.33)
R1 R2
v1 − v+ v+ − vo
= (3.34)
R1 R2
Solving for vo :
  
R1 + R2 R4 R2
vo = v2 − v1 (3.35)
R1 R3 + R4 R1
R2
We can alternatively re-write the first bracket as 1 + R1 and divide the numerator and denominator of
the second bracket by R4 :
  
R2 1 R2
vo = 1+ R3
v2 − v1 (3.36)
R1 R +1
R1
4
R2
1+ R1 R2
= R3
v2 − v1 (3.37)
1+ R4
R1

R2
We can then multiply the numerator and denominator of the first fraction by the reciprocal of R1 . Doing
this will give the gain a nice expression that we can memorise easily:
R1 R2
R2 (1 + R1 ) R2
vo = R1 R3
v2 − v1 (3.38)
R2 (1 + R4 )
R1
Distributing R1 /R2 into the quantity in the numerator, we attain:
R1
R2 +1 R2
vo = R1 R3
v2 − v1 (3.39)
R2 (1 + R4 )
R1
Reciprocating the R1 /R2 in the denominator, we attain the input-output voltage relationship that we
want:
    
R2 1 + R1 /R2 R2
vo = v2 − v1 (3.40)
R1 1 + R3 /R4 R1
3.12 Cascading Op-Amps
We can combine different op-amps in order to achieve a certain output voltage. This process is called
cascading. We can demonstrate the process of cascading op-amps with the following example questions:

3.12.1 Cascade Example 1


Design an op-amp to output:
vo = −3v2 + 5v1 (3.41)
This can be achieved in numerous ways via cascading and is not limited to these following solutions.
However, notice that (3.41) is almost in the form of a difference op-amp output with the signs the other
way around. This gives rise to method 1:

Method 1
Factorising −1 from (3.41), we get:

vo = −(3v2 − 5v1 ) (3.42)


This looks like the output from a difference op-amp that is passed through an inverting op-amp with a
gain of 1. Let’s sort out the difference scenario first.

Consider a difference op-amp with an output of

va = 3v2 − 5v1 (3.43)


Equating coefficients of v1 from (3.43) and (3.40):

R2
= 5 −→ R2 = 5R1 (3.44)
R1
Now, equating coefficients of v2 from (3.43) and (3.40):
  
R2 1 + R1 /R2
=3 (3.45)
R1 1 + R3 /R4
Substituting R2 /R1 = 5 → (3.45):

1 + 51
 
5 =3 (3.46)
1 + R3 /R4
6
=3 (3.47)
1 + R3 /R4
R3
1+ =2 (3.48)
R4

R3 = R4 (3.49)
Using these resistor relationships, the difference op-amp should look something like this:
The output of the difference op-amp can be passed into the inverting terminal of an inverting op-amp.
What this will do is produce an output voltage:

vo = −va (3.50)
= −(3v2 − 5v1 ) (3.51)

The inverting op-amp should have


Ri = Rf
Rf
such that A = Ri = 1. The final circuit used to achieve this looks a little something like this:
Method 2
We can also thing of it as v2 being inverted via an inverting op-amp with a gain of 3 (i.e. va = −3v2 ).

The inverted output will pass into a summing op-amp with a gain of 1 (i.e. needs to pass through a
resistor R1 = Rf ). Meanwhile, another input into the summer receives a gain of 5 (i.e. needs to pass
through a resistor with R2 = 15 Rf )

Stage 1
For the inverting op-amp, the gain needs to be 3. Therefore, we require:

Rf = 3Rin (3.52)
The circuit looks something like this:

Stage 2
The output from the inverted op-amp, va , will pass into the summer along side an input voltage which
will obtain a gain of 5 as it passes through R = 15 Rf . This is shown in the circuit below:
3.13 Differentiator
The differentiator op-amp acts on the input voltage via differentiation to produce an output voltage. The
circuit design for the differentiator is shown below:

The input-output voltage relation for the differentiator is:

dVin
vo = −RC (3.53)
dt

3.13.1 Derivation
The voltage at the non-inverting terminal is v+ = 0. Using the ideal op-amp relationship:

v+ = v− = 0 (3.54)
Hence, performing KCL at the node corresponding to the inverting terminal of the op-amp:

ic (t) = io (3.55)
dq(t) v− −vo
Now, we know that ic (t) = dt = C dvdt
c (t)
. From the diagram, we also know that io = R . However,
because v− = 0:

dvc (t) −vo


C = (3.56)
dt R
However, from the circuit diagram, we can also express the voltage across the capacitor as the potential
difference between the two nodes, vin − v− = vc (t). However, since v− = 0, vin = vc (t). We can hence
re-write (3.56) as:

dvin −vo
=C (3.57)
dt R
Rearranging and making the output voltage the subject of the equation, we obtain the characteristic
equation for the differentiator:

dvin
vo = −RC (3.58)
dt
3.14 Integrator
The integrator uses integration on the input voltage to produce the output voltage. The circuit diagram
for the integrator is shown below:

The characteristic equation describing the integrator op-amp is:

Z t
Vin
vo = vo (0) − dt (3.59)
0 RC

3.15 Derivation
The non-inverting terminal is directly connected to ground. Hence, we can say, along with the ideal
op-amp condition, that:

v+ = v− = 0 (3.60)
Now, performing KCL at the node corresponding to the inverting terminal of the op-amp:
vin − v−
= ic (t) (3.61)
R
Similar to the differentiator, we know that ic (t) = C dvdt
c (t)
. However, from this circuit diagram, we know
that vc (t) = v− − vo . Since v− = 0, then we can say that vc (t) = −vo . We can re-write (3.61) as:

vin −dvo
=C (3.62)
R dt
dvo −vin
= (3.63)
dt RC
integrating both sides from τ = 0 to τ = t, we can write (3.62) as:
Z t
vin
vo (t) − vo (0) = − dτ (3.64)
0 RC
Rearranging, we obtain our required expression:
Z t
vin
vo (t) = vo (0) − dτ (3.65)
0 RC
Chapter 4

Transient Response In First Order RLC


Circuits

The charge or discharge phase of a capacitor does not occur instantaneously. Instead, we describe the
phase as a transient response. There are two cases, one where the capacitors’ voltage undergoes voltage
discharge (decay) or when the capacitor experiences a voltage excitation known as a step-response.

4.1 Transient Behaviour During Discharge In An RC Circuit


Consider the following source-free RC circuit with a capacitor connected in series with voltage source:

The capacitor’s voltage is a function of time. Suppose for t < 0, the capacitors voltage is vc (t < 0) = V0 .
Note, any change to the voltage of the capacitor is NOT instantaneous. Hence, if the discharge proceeds
at t = 0:

vc (0− ) = vc (0+ ) = V0 (4.1)


By KVL, the sum of all voltages around the circuit is 0:

vR (t) − vc (t) = 0 (4.2)


The current through the capacitor, or any element in the circuit in general, is defined as rate of movement
of charge. But for a capacitor, its charge, q is q = Cvc (t). Hence:

dq d(Cvc (t)) dvc (t)


i= = =C (C is a constant) (4.3)
dt dt dt

22
Rearranging (4.2) and incorporating expression (4.3) for the current in the circuit we get the following:

vc (t) = vR (t) (4.4)


= iR (Ohm’s Law) (4.5)
dvc (t) dvc
= −RC (i < 0 since < 0) (4.6)
dt dt
This has become a 1st order ’separable’ ODE with respect to the capacitor’s voltage. Rearranging, we
attain:

dvc (t) dt
− = (4.7)
dt RC
Integrating both sides, we can obtain an equation for the capacitor voltage from t = 0 to t = t

Z t Z t
dvc (t) dt
− = (4.8)
0 dt 0 RC
 
vc (t) t
−ln = (Since vc (0) = V0 ) (4.9)
V0 RC
 
vc (t) t
ln =− (4.10)
V0 RC
vc (t) t
= e− RC (4.11)
V0
Which means the voltage of the capacitor does undergo a discharge in the form of an exponential decay:

t
vc (t) = V0 e− RC

The capacitor’s voltage in this RC circuit scenario can be graphed as a function of time, as shown below.

Note, the expression RC in the equation is called the time-constant, τ :

τ = RC
4.2 Transient Behaviour During Charge Phase In An RC Circuit Due
To Step-Response From Source
This occurs when the RC circuit contains a source that creates a voltage excitation in the circuit and
hence in the capacitor. A switch is usually incorporated in the circuit to achieve this. Observe an example
circuit below:

Figure 4.1: Step-reponse from closing the switch

In order to fully understand the impact of this step-response, the unit-step function denoted by u(t)
must be discussed.

4.2.1 The Unit-Step Function


The unit-step function is given by:


0 t<0
u(t) =
1 t>0

The unit-step function is an example of a singularity (piecewise) function, functions which exhibit dis-
continuity or non-differentiability at certain values.

Note that that the unit-step function is undefined at t = 0 since the change from 0 to 1 is abrupt. The
change in the unit-step function need not be always at t = 0 (though it is convenient that way) and could
occur at some t = t0 , which in this case would mean:

0 t < t0
u(t = t0 ) =
1 t > t0

Likewise, if the change occurred at some time t = −t0 , we would write the unit-step function in the same
fashion:

0 t < −t0
u(t = t0 ) =
1 t > −t0
The graph of the unit-step function is shown below where it is clear of the non-differentiability at t = 0:

4.2.2 Role Of Unit-Step In RC Step-Response


The unit-step function is used to algebraically indicate the voltage excitation that occurs when the switch
is closed in Fig 4.1. Moreover, it appears in the final equation for the capacitor voltage in this scenario.

Derivation
Observe Fig 4.1(a). At t < 0, let the capacitor voltage be V0 . Any change to the voltage will not be
instantaneous. Hence:

vc (0− ) = vc (0+ ) = V0 (4.12)


Now, observe Fig 4.1(b). For t ≥ 0, the entire circuit experiences a voltage excitation due to the voltage
source upon circuit completion. Because the source experiences an excitation, the unit-step function
comes in:

Vs −→ Vs u(t) (4.13)
Performing KVL around the circuit loop in Fig 4.1(b), the sum of all voltages of all elements in the circuit
must add to 0. Hence:

vR (t) + vc (t) = Vs u(t) (4.14)


i(t)R + vc (t) = Vs u(t) (4.15)

Similar to the source-free RC circuit, the current through all elements is the same. Then, the current can
be expressed as:

dq d(Cvc (t)) dvc (t)


i(t) = = =C (C is a constant) (4.16)
dt dt dt
Hence, we modify (4.16) which gives us:

dvc (t)
RC + vc (t) = Vs u(t) (4.17)
dt
Now, for t > 0, u(t) = 1 (from the unit-step function). Hence, the equation becomes:

dvc (t)
RC + vc (t) = Vs (4.18)
dt
This is, once again, a separable 1st order ODE. Upon rearranging the equation, we obtain an integrable
equation:
dvc (t) dt
= (4.19)
Vs − vc (t) RC
Multiplying both sides by −1 to ease the integration:

dvc (t) dt
=− (4.20)
vc (t) − Vs RC
We then integrate both sides both sides with respect to time from t = 0 −→ t

Z t Z t
dvc (t) dt
= − (4.21)
vc (t) − Vs RC
0  0
vc (t) − Vs t
ln =− (since vc (0) = V0 ) (4.22)
V0 − Vs RC

Making both sides as powers of e, the ln is canceled which gives us:

vc (t) − Vs t
= e− RC (4.23)
V0 − Vs
Making the vc (t) the subject, we obtain an equation for the voltage of the capacitor in the context of a
step-response from the voltage source.

vc (t) = Vs + (V0 − Vs )e−t/RC

Including the unit-step-function in the equation gives us:

vc (t) = Vs u(t) + (V0 − Vs u(t))e−t/RC


This means like the unit-step function, the capacitor voltage can be written as a piecewise function as
well. Note that for t < 0, vc (t < 0) = V0 . Hence:


V0 t<0
vc (t) = t
Vs + (V0 − Vs )e− RC t>0

If the capacitor was initially uncharged, then V0 = 0. The function for the capacitor’s voltage then be-
comes:

When V0 = 0:

0 t<0
vc (t) = t
Vs (1 − e− RC ) t>0
4.3 Short-Cut Equation
Observe again the equation:

vc (t) = Vs + (V0 − Vs )e−t/RC

The capacitor voltage will follow this behaviour for t > 0. After a long period of time, the capacitor will
equal to the step-voltage that caused the voltage excitation. i.e.

Vs = vc (∞) (4.24)
Moreover, we know that V0 is the capacitor voltage for t < 0. However, we also know that its voltage
does not change instantaneously when the step-response is made. Hence:

V0 = vc (0) (4.25)
Hence, our short-cut equation becomes:

Short-Cut Equation
vc (t) = vc (∞) + vc (0) − vc (∞) e−t/RC
 

4.4 Finding vc (∞) and vc (0) via Example


In most scenarios, the step response occurs due to a change in the position of a switch. The circuit will
contain more than one resistor and voltage division will occur. Take for example this circuit:

For this situation, we want to determine the capacitor’s voltage as a function. We firstly need to determine
vc (0− ) when the switch was opened:

For vc (0), provided the switch was opened for a LONG time before closing:

• The capacitor was fully charged due to the 15V source.


• Hence, vc (0) = 15V
When the switch closes at t = 0, we note again that the capacitor voltage does NOT change instanta-
neously. Hence:

vc (0− ) = vc (0+ ) = vc (0) = 15V (4.26)

Much after t = 0, the circuit behaves like the following:

When t −→ ∞ (i.e. t > 0), the capacitor is fully charged. At this point, the capacitor acts like a
short-circuit (no current flows to it). Hence, we can perform KCL to the middle node:

15 − vc (∞) vc (∞) − (−7.5)


= (4.27)
2 6

Solving for vc (∞), we find that:

vc (∞) = 9.375V (4.28)


The last thing we need to do is to determine the Thevenin resistance of the circuit with respect to the
terminals of the capacitor. Let the positive terminal be a and the negative terminal be b. Then turn off
all independent sources:

From the circuit diagram, RT h = 23 Ω. It was given that C = 13 F . Hence, the time-constant, τ for the
circuit is:
  
3 1 1
τ = RC = = (4.29)
2 3 2
The voltage of the capacitor is hence:

vc (t) = vc (∞) + vc (0) − vc (∞) e−t/τ


 
(4.30)
= 9.375 + 15 − 9.375 e−t/0.5
 
(4.31)
= 9.375 + 5.625e−2t (4.32)
4.5 Summary Of Steps To Determine Capacitor Voltage From Step-
Response
To summarise:

1. Observe position of switch for t < 0:

(a) If open −→ open-circuit.


(b) If closed −→ short-circuit.

2. Determine vc (0) by either:

(a) Using voltage division over a resistor that shares the same voltage.
(b) Using techniques such as KVL or KCL to solve for the node voltage at the capacitor.

3. Observe the position of the switch when t ≥ 0. Note that at t = 0, the voltage of the capacitor
remains the same as its voltage does NOT change instantaneously. vc (0− ) = vc (0+ ) = V0 .


4. Perform KVL/KCL or voltage division in order to determine the voltage across the capacitor when
t > 0 (i.e. vc (∞)).

(a) If using KCL, note that the capacitor acts as an open-circuit i.e. it acts as an infinite resis-
tance. Hence, no current flows across capacitor.

5. Determine the Thevenin resistance (RT h ):

• If only independent sources, remove all sources and then determine Thevenin resistance.

6. Determine the time-constant, τ by τ = RC (C is given).

7. Substitute vc (0), vc (∞) and τ into the short-cut equation.


4.6 Transient Behaviour During Discharge In An RL Circuit
An RL circuit, like an RC variant has resistors and an inductor in the circuit. For instance, observe the
circuit below:

Figure 4.2: Source-free RL circuit

When working with RL circuits, we consider the ODE in terms of the inductor current rather than voltage
in the RC scenario. We consider two scenarios; When t < 0 and when t ≥ 0. Assume that:

iL (t < 0) = iL (0− ) = iL (0+ ) = I0 (4.33)


We use the fact that the inductor cannot change abruptly. Applying KVL to the circuit loop:

vL (t) + vR (t) = 0 (4.34)


vL (t) = −vR (t) (4.35)

We know that the voltage drop across the capacitor is proportional to the rate of change of current
through the inductor:

di(t)
vL (t) = L (4.36)
dt
And by Ohm’s Law, we know the current through the resistor is given by:

vR (t) = i(t)R (4.37)


Substituting into (4.36) gives us a 1st order separable ODE that we can rearrange and integrate:

di(t)
L = −i(t)R (4.38)
dt
di(t) R
= − dt (4.39)
i(t) L
Z t Z t
di(t) R
= − dt (4.40)
0 i(t) 0 L
 
i(t) R
ln = − t (i(0) = I0 ) (4.41)
I0 L
And making i(t) the subject, we find that the current, due to the presence of the inductor behaves as an
exponential decay:

t
i(t) = I0 e− τ
L
Where τ = R
4.6.1 Voltage across the inductor in a source-free RL circuit
Now that we have an equation for the current through the inductor, we can obtain an equation for the
voltage across the inductor. Recall:

di(t)
vL (t) = L (4.42)
dt 
d −R t
=L I0 e L (4.43)
dt
  
R −R t
=L − I0 e L (4.44)
L

Hence, the voltage across an inductor in a source-free circuit is given by:

t
vL (t) = −I0 Re− τ

4.6.2 Voltage across the resistor in a source-free RL circuit


Using Ohm’s Law, we find that:

vR (t) = i(t)R (4.45)


− τt
= I0 Re (4.46)

t
vR (t) = I0 Re− τ
4.7 Transient Behaviour During Charge Phase In An RL Circuit Due
To Source Excitation
Upon source excitation, the inductor undergoes a charge phase, like an RC circuit, that arises due to the
position change of a switch. Observe an RL circuit for t < 0 and t ≥ 0, respectively:

Consider the left circuit (t < 0). Assume that the inductor already had an initial current passing through,
I0 . Upon any source excitation, the current cannot change abruptly:

i(t < 0) = i(0− ) = i(0+ ) = I0 (4.47)


Now, when t = 0, switch closes. The charge phase and transient behaviour occur when t > 0 in the right
circuit. Performing KVL in the right-circuit, we find:

vR (t) + vL (t) = Vs u(t) (4.48)


Where u(t) is the unit-step function. When t > 0, u(t > 0) = 1. Hence:

vR (t) + vL (t) = Vs (4.49)


Since the voltages on the LHS can be expressed in terms of the circuit current, we can express the equation
as a first order ODE in terms of the circuit current, i(t):

di(t)
i(t)R + L = Vs (4.50)
dt
di(t) Vs R
= − i(t) (4.51)
dt L L 
1
= Vs − Ri(t) (4.52)
L
(4.53)
Rearranging, we can form the following integrable equation:

di(t) dt
= (4.54)
Vs − Ri(t) L
Multiplying both sides by -1 to make the equation easier to handle, we find:

di(t) dt
=− (4.55)
Ri(t) − Vs L
Z t Z t
1 Rdi(t) dt
= − (4.56)
R 0 Ri(t) − Vs 0 L
(4.57)
 
Ri(t) − Vs R
ln =− t (4.58)
RI0 − Vs L
Ri(t) − Vs R
= e− L t (4.59)
RI0 − Vs
Solving for i(t), we now have a function for the circuit current as a function of time due to the presence
of the inductor:

 
Vs Vs −t/τ
i(t) = + I0 − e
R R

The expression VRs is current after a long time (t > 0) the position change of the switch. I0 corresponds
to the starting current. Like the RC, we can say:

I0 = i(0); (4.60)
Vs
= i(∞) (4.61)
R
With this, we can make a short-cut equation that makes any RL circuit a lot easier:

Short-Cut Equation:

i(t) = i(∞) + i(0) − i(∞) e−t/τ


 

4.8 Finding i(0) and i(∞) via Example


An example of an RL circuit is shown below:

Solving For i(0):


For a long time during t < 0. The switch was closed. Therefore, no current will flow into the 5Ω resistor
(since current will flow through the path of least resistance). Moreover, for very long periods, the
inductor acts a short-circuit. We can reduce the circuit for t < 0 into:
The circuit current will indeed be 6A. To further justify this, a source transform can be made on the 6A
source and the 10Ω resistor:

For a long period of time during t < 0, the inductor acts as a short-circuit. Hence, we can find i(0) simply
from Ohm’s Law:

Vs 60
i(0) = = = 6A (4.62)
R 10

Solving for i(∞):


When t = 0, the switch is opened. Hence, for t > 0, the open-switch acts as an open-circuit with infinite
resistance. Our circuit during this period is shown below:

To current to be determined is shown in the circuit as i. This is the i(∞) to be solved for. An easier way
to solve for i would be to first perform a source transformation on the circuit:

Again, for a long time after t = 0, the inductor acts a short-circuit. Hence, i(∞) can be solved for via
Ohm’s Law:
Vs 60
i(∞) = = = 4A (4.63)
RT h 5 + 10
The value for Rt h is found by switch off the voltage source (which becomes a short-circuit) and again
treating the inductor as a short as well. Hence, the resistors are in series and add linearly.

RT h = 15Ω (4.64)
We can now find the time-constant, τ :

L 15 1
τ= = = seconds = 0.1seconds (4.65)
R 1.5 10
Recalling our short-cut equation for the transient behaviour of the circuit current due to the inductor:

i(t) = i(∞) + (i(0) − i(∞))e−t/τ (4.66)


Upon substitution and simplification we conclude that:

i(t) = 4 + (6 − 4)e−t/(0.1) (4.67)


−10t
= (4 + 2e ) [A] (4.68)
4.9 Summary Of Steps To Determine Circuit (Inductor) Current From
Step-Response
The procedure to determine the circuit (inductor) current is much the same as that to find the capacitor
voltage with a few minor changes.

To summarise:

1. Observe the position of the switch at t < 0:

• If switch is open −→ open-circuit (no current will flow in that circuit segment)
• If switch is closed −→ short-circuit (current will flow in the segment of the circuit)

2. The position of the switch will determine the circuit for t < 0. Use any technique such as KCL,
KVL, current/voltage division to determine the circuit current, i(0) during this time period. Source
transformation can be use full to convert current sources into voltage sources.

• The inductor, for a long period of time, will act as a short in the circuit. i.e. it plays no role
in i(0)

3. Observe the position change of the switch at t = 0:

• The new circuit will be determined by the change in position of the switch.

4. Use KCL, KVL, voltage/current divider and/or source transform to determine the circuit current
i(∞) during t > 0.

5. In the circuit for t > 0, find the Thevenin resistance, RT h .

• Switch off all independent sources and treat the inductor, again, as a short in the circuit.
L
6. Then, given the value of L, calculate the time-constant via the equation τ = R

7. Lastly, substitute i(0), i(∞) and τ into the short-cut equation: i(t) = i(∞) + i(0) − i(∞) e−t/τ
 
4.10 RL-Analysis Using Calculus
Sometimes, the calculus approach may be an easier alternative. This will simply rely on mesh analysis or
nodal analysis. Simultaneous equations can be solved and will result in a first order ODE which can be
solved for something like current.

Examples
Find i and vx in the circuit of Fig. 7.15. Let i(0) = 12A.

Solution
Here is when calculus will prove useful since there is no switch scenario. Consider the circuit consisting
of two mesh currents i1 (left loop) and i2 (right loop).

First Loop

1i1 + 2(i1 − i2 ) + 2vx + vL (t) = 0 (4.69)


di1 (t)
3i1 − 2i2 + 2vx = −L (4.70)
dt
However, we know that vx = i1 × 1Ω = i1 . Upon substitution and simplification, we attain:

di1 (t)
5i1 − 2i2 = −L (4.71)
dt
Second Loop

6i2 − 2vx + 2(i2 − i1 ) = 0 (4.72)


−2i1 + 2i2 − 2vx = 0 (4.73)

However, we know that vx = i1 × 1Ω = i1 . Upon substitution and simplification, we attain:

1
i2 = i1 (4.74)
2
Substitute (4.75) −→ (4.72)
 
1 di1
5i1 − 2 i1 = −L (4.75)
2 dt
di1
4i1 = −2 (L = 2H) (4.76)
dt
di1 (t)
2i1 (t) = − (L = 2H) (4.77)
dt
We can then re-arrange the equation to form an integrable equation:

di1 (t)
−2dt = (4.78)
dt
Z t Z t
di1 (τ )
−2 dτ = (4.79)
0 0 dτ
 
i(t)
−2t = ln (Note: i1 = i) (4.80)
i(0)
(4.81)

But we were given that i(0) = 12A. Hence:

i(t)
e−2t = (4.82)
i(0)
i(t) = 12e−2t [A] (4.83)

For the voltage across the 1Ω resistor, we use Ohm’s law. Note that from the direction of the current, i,
which we have defined as positive, this means that by the passive sign convention, the voltage across the
1Ω resistor must be negative:

vx (t) = −i × 1Ω (4.84)
= −i (4.85)
= −12e−2t [V ] (4.86)
(4.87)
Chapter 5

Transient Response In Second Order


RLC Circuits

As with the 1st order RLC circuits, the source-free and step-response scenarios will be derived. These
will involve 2nd order differential equations that exhibit a characteristic equation.

5.1 Source-Free RLC Series Circuit:


Observe the following circuit containing a resistor, inductor and capacitor in series:

Since this is a circuit with all elements in series, it is more meaningful to derive an expression for the
current that passes through all elements rather than the voltage drop across each element. Let the current
be i(t). From KVL:

iR + vL (t) + vC (t) = 0 (5.1)


Expressing each term in terms of i:
Z t
di 1
iR + L + vC (0) + i(τ )dτ = 0 (5.2)
dt C 0
Since this is a source-free RLC circuit, the voltage across the capacitor would be 0 for t = 0 (unless
otherwise stated). Hence:

1 t
Z
di
iR + L + i(τ )dτ = 0 (5.3)
dt C 0
We now differentiate both sides with respect to time. By the 1st Fundamental Theorem of Calculus, this
will eliminate the integral and will allow us to solve the ODE:

1 d t
    Z
d d di
iR + L + i(τ )dτ = 0 (5.4)
dt dt dt C dt 0

39
Hence, we arrive at:

d2 i di 1
L 2
+R + i=0 (5.5)
dt dt C

d2 i R di 1
+ + i=0 (5.6)
dt2 L dt LC
di di
This a 2nd order homogeneous ODE in terms of dt . We can solve this by setting λ = dt :

R 2 1
λ2 + λ + =0 (5.7)
L LC
Using the quadratic formula, we find that the roots can be expressed as:
q
−RL ± (R 2 1
L ) − 4(1)( LC )
λ1,2 = (5.8)
2
s 
R 2
 2
−R 1
λ1,2 = ± − √ (5.9)
2L 2L LC
We can write the roots in a much simpler form:
q
λ1,2 = −α ± α2 − ω02 (5.10)
The quantities, α and ω0 are given the names:

Undamped (Neper) Natural Frequency:

R
α= (5.11)
2L
Damping Factor:
1
ω0 = √ (5.12)
LC

Given that we know α and ω0 , we have three types of solutions to this 2nd order ODE. These types are
dependent on the discriminant, of the roots.
5.1.1 Overdamped Case (α > ω0 )
i.e. this is when α2 − ω02 > 0. This means there are two distinct roots to the ODE. Hence, the current
can be expressed as:

i(t) = Aeλ1 t + Beλ2 t (5.13)

Where A, B ∈ R. These constants can be evaluated given initial conditions. The current as a function of
time can be shown in a graph:

5.1.2 Critically-Damped (α = ω0 )
i.e. This is when α2 − ω02 = 0. This means the root is a double root of the ODE. Hence, the current can
be expressed as:

i(t) = Ceλ1 t + Dteλ2 t (5.14)

Where C, D ∈ R. The current as a function of time can be shown in a graph:

5.1.3 Underdamped Response (α < ω0 )


i.e. this is when α2 − ω02 < 0. This means the roots of the ODE are complex and can be written in the
form:

λ1,2 = −α ± jωd (5.15)


≡ a ± jb (5.16)
Associated with the underdamped case is the damped natural frequency or damping frequency. This is
denoted by ωd as shown in (5.15). This frequency can be defined as follows:

q
λ1,2 = −α ± α2 − ω02 (5.17)

However, α2 − ω02 < 0. This means we can re-write (5.17) as the following:

q
λ1,2 = α ± −(ω02 − α2 ) (5.18)
q
= α ± j ω02 − α2 (5.19)
= α ± jωd (5.20)

q
ωd = ω02 − α2 (5.21)

Because the roots are complex, the current can be expressed in the form:
 
at
i(t) = e Ecos(bt) + F sin(bt) (5.22)

 
−αt
i(t) = e Ecos(ωd t) + F sin(ωd t) (5.23)

The current as a function of time is shown below. Note that the current oscillates within the exponential
asymptotes. The amplitude of the oscillation decreases as t → ∞:
5.2 Source-Free RLC Parallel Circuit
Consider the presence of the same elements but arranged in a parallel circuit:

Because we are dealing with a parallel configuration, it is more meaningful to talk about the voltage drop
across each element. Performing KCL at the top node:
v−0
+ iL (t) + iC (t) (5.24)
R
We would like to express all terms in (5.20) in terms of voltages:

1 t
Z
v dv(t)
+ i(0) + v(τ )dτ + C =0 (5.25)
R L 0 dt
Let us suppose that at t = 0, i(0) = I0 where I0 ∈ R. Hence, we can re-write (5.21) as:

1 t
Z
v dv(t)
+ I0 + v(τ )dτ + C =0 (5.26)
R L 0 dt
Differentiating both sides of (5.22) will eliminate the integral via the 1st Fundamental Theorem of Calculus.
Hence:

1 d t
    Z  
d v d d dv(t)
+ I0 + v(τ )dτ + C =0 (5.27)
dt R dt L dt 0 dt dt

1 dv(t) 1 d2 v(t)
+ v(t) + C =0 (5.28)
R dt L dt2

Rearranging the equation a bit, we attain a 2nd order ODE in terms of v(t):

d2 v(t) 1 dv(t) 1
+ + v(t) = 0 (5.29)
dt RC dt LC
dv
To solve this ODE, set s = dt :

1 1
s2 +s+ =0 (5.30)
RC LC
Using the quadratic formula, we find the roots of the quadratic to be:

s 2  2
1 1 1
s1,2 = − ± − √ (5.31)
2RC 2RC LC
q
= −α ± α2 − ω02 (5.32)
We can see that the expressions for the neper and undamped frequency have changed slightly given a
parallel configuration:

Neper Frequency:
1
α= (5.33)
2RC
Undamped Natural Frequency:
1
ω0 = √ (5.34)
LC

And as per usual:

Damped Natural Frequency: q


ωd = ω02 − α2 (5.35)
If and only if the circuit system results in an underdamped response

The waveforms for the voltage response in the parallel circuit are in the same form as the current functions
for the RLC series circuit:

Overdamped (α > ω0 ):
v(t) = Aes1 t + Bes2 t (5.36)
Critically-damped (α = ω0 ):
v(t) = Ces1 t + Dtes2 t (5.37)
Underdamped (α < ω0 ):
 
−αt
v(t) = e Ecos(ωd t) + F sin(ωd t) (5.38)
5.3 Step-Response In An RLC Circuit + Voltage Source
In both cases, either a voltage or current source is added in a series or a parallel circuit, respectively. The
difference is essentially an extra term in the solution for current or voltage waveforms:

5.3.1 Step-Response In Series RLC Circuit

Figure 5.1: Series RLC Circuit

Due to the presence of a DC voltage source, we would like to attain a characteristic equation in terms of
voltage, specifically v(t). KVL around the loop gives:

di
iR + L
+ v(t) = Vs (5.39)
dt
However, we know the current through the capacitor, and hence through all elements is:

dv
i=C (5.40)
dt
Substituting → (5.35):

dv d2 v
RC + LC 2 + v(t) = Vs (5.41)
dt dt
Simplification gives our usual second-order ODE but this time, non-homogenous:

d2 v R dv v(t) Vs
2
+ + = (5.42)
dt L dt LC LC
The non-homogeneity of the ODE and the presence of a DC source gives rise to a waveform of the form:

v(t) = vsteady−state (t) + vtransient (t) (5.43)


vtransient (t) retains equations in the same form as that for the source-free series RLC scenario. The
steady-state voltage is simply the voltage source responsible for the step-response. Hence for the three
cases, we have:

Overdamped (α > ω0 ):
v(t) = Vs + Aes1 t + Bes2 t (5.44)
Critically-damped (α = ω0 ):

v(t) = Vs + Ces1 t + Dtes2 t (5.45)

Underdamped (α < ω0 ):
 
−αt
v(t) = Vs + e Ecos(ωd t) + F sin(ωd t) (5.46)
5.3.2 Step-Response In A Parallel RLC Circuit + Current Source

Figure 5.2: Step-response for a parallel RLC circuit with a current source.

The presence of the current source suggests that our characteristic equation should be in terms of current,
specifically, the current through the inductor, i(t). For t > 0, KCL at the top node gives:

Is = iR + i(t) + ic (5.47)
However, by Ohm’s Law, and the equations describing the capacitor and inductor, we know that:
 
vc vL L di
iR = = = (Because R k C and L k C) (5.48)
R R R dt

dvc dvL
ic = C =C (Because L k C) (5.49)
dt dt  
d di
= LC (5.50)
dt dt
d2 i
= LC 2 (5.51)
dt
Substituting these expressions into (5.43), we attain:

d2 i L di
 
LC 2 + + i(t) = Is (5.52)
dt R dt
A bit of simplification gives us a second-order ODE in terms of the current through the inductor:

d2 i
 
1 di 1 Is
+ + i(t) = (5.53)
dt2 RC dt LC LC
When solving for i(t), the equations are in the form of the equations describing the source-free parallel
RLC circuit. Hence, the equations are:

Overdamped (α > ω0 ):
i(t) = Is + Aes1 t + Bes2 t (5.54)
Critically-damped (α = ω0 ):
i(t) = Is + Ces1 t + Dtes2 t (5.55)
Underdamped (α < ω0 ):
 
−αt
i(t) = Is + e Ecos(ωd t) + F sin(ωd t) (5.56)
Chapter 6

Sinusoidal Signals, Sinusoidal


Steady-State Analysis Of Circuits

47
Chapter 7

AC Power Analysis

In AC circuits, unlike DC circuits, voltage, current and resistance/impedance are not necessarily constant.
The input of an AC source means all these quantities exhibit periodic behaviour in the form of sinusoids.
This impacts on our understanding of power in AC. From here, we need to take into account that power
can take different forms due to these sinusoids.

7.1 Instantaneous Power


Instantaneous power is the power as any instant of time, t. In the DC analogue, P = V I. In AC, we
express power, instantaneously, as:

p(t) = v(t) · i(t) (7.1)


Where the instantaneous voltage and current take the form of sinusoidal functions with a phase an-
gle/offset of θv and θi respectively:

v(t) = Vm cos(ωt + θv ) (7.2)


i(t) = Im cos(ωt + θi ) (7.3)

Substituting (7.2) and (7.3) into (7.1), we get:

p(t) = Vm Im · cos(ωt + θv ) · cos(ωt + θi ) (7.4)


Now, consider the following trigonometric identity:
 
1
cosA · cosB = cos(A + B) + cos(A − B) (7.5)
2
Let A = ωt + θv and B = ωt + θi . Substituting into (7.5) and multiplying Vm Im to both sides, we get:
 
1
Vm Im cos(ωt + θv ) · cos(ωt + θi ) = Vm Im cos(2ωt + θv + θi ) + cos(θv − θi ) (7.6)
2
In other words, this is an alternate expression for instantaneous power:

1 1
p(t) = Vm Im cos(2ωt + θv + θi ) + Vm Im cos(θv − θi ) (7.7)
2 2

48
7.2 Average Power
It is very difficult to measure instantaneous power. Instead, the average power of a system is an easier
quantity to measure. The average value of a sine wave is the area under the curve for one period of the
sine function:
Z T
1
Average Value = f (t)dt (7.8)
T 0
In the context of average power, this would be the average value of the instantaneous power function. Let
P = Average Power:
Z T
1
P = p(t)dt (7.9)
T 0
Substituting (7.7) into the integral (7.9), we get:

1 T 1
Z  
1
P = Vm Im cos(2ωt + θv + θi ) + Vm Im cos(θv − θi ) dt (7.10)
T 0 2 2
The integral of a sum is the linear sum of two integrals. Hence, we can write (7.10) as:
Z T Z T
1 1 1 1
P = Vm Im cos(2ωt + θv + θi )dt + Vm Im cos(θv − θi )dt (7.11)
T 0 2 T 0 2
Now, immediately note that integral of any time-independent sinusoidal function (regardless of phase
offset) will always be equal to 0. Hence, the first integral in (7.11) becomes 0:
Z T
1 1
P =0+ Vm Im cos(θv − θi )dt (7.12)
T 0 2
The reason we can reduce the integral is due to the symmetry of the sine function. Integrating over a
period means the sum of two oppositely signed areas which results in 0. This is graphically shown below:

Now, we note in (7.12) that the integrand is time-independent which means it is not affected by the
integral. Hence, it is a constant:

Z T
1 1
P = Vm Im cos(θv − θi ) dt (7.13)
2 T 0
 T
1 1
= Vm Im cos(θv − θi ) t (7.14)
2 T 0
1 1
= Vm Im cos(θv − θi ) T (7.15)
2 T
Hence, the average power of a linear AC circuit system is given by:

1
P = Vm Im cos(θv − θi ) (7.16)
2

7.3 Average Power (Frequency Domain)


We begin the average power in the time-domain:
1
P = Vm Im cos(θv − θi ) (7.17)
2
We can express P in the frequency domain (phasor) by first turning (7.2) and (7.3) into phasors:

V = V m θv (7.18)
I = Im θi (7.19)

Now, in order to get the phase difference, θv − θi but with phasors, we can take the conjugate of (7.19):

I∗ = Im −θi (7.20)

Multiplying V and I∗ , we attain:

VI∗ = Vm Im θv − θi (7.21)
   
= Vm Im cos(θv − θi ) + j Vm Im cos(θv − θi ) (7.22)

Taking the real part of both sides:

Re[VI∗ ] = Vm Im cos(θv − θi ) (7.23)

Multiplying 12 to both sides, notice that the RHS of (7.23) becomes the same expression as that for the av-
erage power in (7.17). Hence, the average power absorbed by a load (even if the load has a complex value):

1 1
P = Re[VI∗ ] = Vm Im cos(θv − θi ) (7.24)
2 2

7.3.1 Special Cases


Equation (7.24) would be used mainly to calculate the average power of a certain element in an AC circuit
that we consider to be the load of the circuit. Depending on the load, we can simplify our AC analysis
a great deal:

1. Load is a capacitor/inductor: i.e load is purely reactive. Because of this, θv − θi = ±90◦ . If


the phase difference is a multiple of 90◦ , cos(θv − θi ) = 0. Hence:

PC = PL = 0 (7.25)
2. Load is a resistor: i.e. load is purely resistive. In this case, the phase angle of the voltage
across the resistor and the phase angle of the current are equal. i.e. θv − θi = 0. This means (7.24)
will be non-zero. Hence, we can express the average power of a resistive load as:
1 1 1 2 1
P = Re[VI∗ ] = Vm Im = Im R = |I|2 R (7.26)
2 2 2 2
Where possible, use P = |I|2 R (a lot easier).

To summarise, we can say:

General Load: P = Vm Im cos(θv − θi )


Purely Reactive Load: PC = PL = 0
Purely Resistive Load: P = 12 |I|2 R
7.4 Maximum Average Power Transfer In AC
The process behind finding the maximum average power transfer in the AC analogue is similar to the
DC scenario with the inclusion of Thevenin impedance as well as Thevenin resistance. Consider the
following Thevenin equivalent circuit:

Where in the most general case, the impedances shown can be represented as:

Zth = Rth + jXth (7.27)


ZL = RL + jXL (7.28)

The condition for maximum average power transfer is:

ZL = Z∗th (7.29)
Note that the load impedance has to equal the conjugate of the Thevenin impedance.

Hence, the condition for maximum average power transfer in an AC circuit is given by:

|Vth |2
Pmax = (7.30)
8Rth

If the load of our concern in purely real (i.e. the load is simply a resistor), then the condition for
maximum power is given by:

RL = |Zth | (7.31)
7.4.1 Example Of Maximum Average Power Transfer Analysis
Consider the following AC circuit:

Step 1: Finding The Thevenin Impedance


Just like in the DC analogue, we would short any independent voltage source and open any independent
current sources. Since the only independent source is the 10∠0◦ V source, we short that and find Zth :

4(8 − 6j)
Zth = 5j + (7.32)
4 + 8 − 6j
 
44 8
= − j Ω (7.33)
15 15

Step 2: Find the Thevenin (Open-Circuit) Voltage


We bring back the original circuit and remove the load impedance. We then find the open-circuit voltage
at the terminals where the load once was. We can use voltage division across the 8 − 6j impedance:

 
8 − 6j 11 2
Vth = V8−6j = (10) = − j (7.34)
4 + 8 − 6j 15 15
= (7.454∠−10.3◦ )V (7.35)
Calculating Maximum Power Transfer
Hence, we can now calculate the maximum power transfer:

|Vth |2
Pmax = (7.36)
8Rth
(7.454)2
= (7.37)
8 · Re[Zth ]
(7.454)2
= 44 (7.38)
8( 15 )
= 2.368W (7.39)
7.5 Effective/RMS Value

The effective or RMS (root mean square) value of a periodic current/voltage is the DC cur-
rent/voltage that delivers the same average power to a resistor as a periodic current/voltage.

i.e. We want to have a way to compare the current/voltage supplied by an AC source and the cur-
rent/voltage supplied by a DC source. To do this, we consider the power supplied by each scenario:

DC Power
DC power can be calculated by:

2
P = Ieff R (7.40)
Where Ieff is the effective/RMS current DC power.

AC Power
The average power, as derived in the first section of this chapter, is given as:
Z T
1
P = i2 Rdt (7.41)
T 0
i.e. since R is constant, we can take it out of the integral, which gives us:
Z T
R
P = i2 dt (7.42)
T 0

RMS Current
Hence, in order to express an the power delivered by an AC source in a DC analogue, we compare equation
(7.40) and (7.42):

R T 2
Z
2
Ieff R = i dt (7.43)
T 0

1 T 2
Z
2
Ieff = i dt (7.44)
T 0
Hence, taking the square root of both sides, we find the effective current to be:
s
1 T 2
Z
ieff = i dt (7.45)
T 0
The expression on the RHS, in general, is defined to be the root-mean-square (RMS) of a periodic func-
tion. Hence, we say that:

s
Z T
1
IRMS = i2 dt (7.46)
T 0
Similarly for an AC voltage source with voltage output v(t), we can say that:

s
Z T
1
VRMS = v 2 dt (7.47)
T 0

7.5.1 RMS-Amplitude Relationship


Suppose we have a periodic function for current, with amplitude Im , in the form:

i(t) = Im cos(ωt) (7.48)


To find the RMS current, we use (7.46) to do so:
s
1 T 2
Z
IRM S = I cos2 (ωt)dt (7.49)
T 0 m
Using double-angle identities, we can say cos2 (ωt) = 12 (1 + cos(2ωt)). Substituting in (7.49), we obtain:

s
2
Z T
Im
IRM S = (1 + cos(2ωt))dt (7.50)
2T 0
s T
2

Im 1
= t + sin(2ωt) (7.51)
2T 2 0

Substituting t = T into the sine function makes it equal to 0 as mentioned in section 7.2. Hence:

s
2
 
Im
IRM S = (T + 0) − (0 + 0) (7.52)
2T
r
2
Im
= ·T (7.53)
2T
(7.54)

Im
IRM S = √ (7.55)
2

And similarly for a periodic voltage input:

Vm
VRM S = √ (7.56)
2
7.6 Example RMS Question
Find the rms value of the current waveform shown below. If the current flows through a 9Ω resistor,
calculate the average power absorbed by the resistor.

Solution
The current waveform over one period (e.g. 0 ≤ t ≤ 2) is given by:
(
16t 0≤t≤1
i(t) =
−16t + 32 1 ≤ t ≤ 2

The RMS current is given by:


s
Z T
1
IRMS = i2 dt (7.58)
T 0

Because the waveform is piecewise, the integral can be split into two different integrals over the two
different periods:

s Z
1 Z 2 
1
IRMS = (16t)2 dt + (−16t + 32)2 dt (7.59)
2 0 1
s  
1 256 256
= + (7.60)
2 3 3
= 9.2376A (7.61)

Hence, if the current flows through a 9Ω resistor, we can say that:

2
PAve = IRMS R (7.62)
2
= (9.2376) (9) (7.63)
= 27.3541W (7.64)
7.7 Apparent Power & Power Factor

The Apparent Power (unit: VA or Volt-Ampere) is the product of the RMS current and
voltage.

As we have seen from the first section, the average power was determined to be:
1
P = Vm Im cos(θv − θi ) (7.65)
2
V Im
However, because we know that VRM S = √m
2
and IRM S = √
2
. We can express (7.65) alternatively as:

Vm Im
P = √ √ cos(θv − θi ) (7.66)
2 2
= VRMS IRMS cos(θv − θi ) (7.67)
= Scos(θv − θi ) (7.68)

Mathematically, the Apparent Power is expressed as:

S = VRMS IRMS (7.69)

The Power Factor is the cosine of the phase different between the current and voltage angles
and gives rise to the real power:
P
pf = cos(θv − θi ) = (7.70)
S

Suppose the circuit has a load impedance, Z. If the voltage across the load is V and the current through
the load is I, then:


V VRM S / 2
Z= = √ (7.71)
I IRM S / 2
VRM S
= (7.72)
IRM S
VRM S
= ∠θv − θi (7.73)
IRM S

The Power Factor is also the cosine of the angle of the load impedance.

7.7.1 What Leading and Lagging Mean


• If θv < θi =⇒ Current leads voltage

• If θv > θi =⇒ Current lags voltage


7.8 Example Question Of Apparent Power & Power Factor
7.9 Complex Power

The Complex Power, S (in VA) is defined to be:


1
S = VI∗ (7.74)
2
Where V and I are phasors for voltage and current and I∗ is the complex conjugate of the phasor
current.

We can express the complex power in terms RMS values in the following way:

1
S = VI∗ (7.75)
2
1
= V I∠θv − θi (7.76)
2
V I
= √ √ (cos(θv − θi ) + jsin(θv − θi )) (7.77)
2 2
= VRMS IRMS (cos(θv − θi ) + jsin(θv − θi )) (7.78)
= VRMS IRMS cos(θv − θi ) + jVRMS IRMS sin(θv − θi ) (7.79)
= P + jQ (7.80)

Complex Power can be defined in rectangular form in terms of RMS values. Equating (7.79)
and (7.80), we attain:

P = VRMS IRMS cos(θv − θi ) (Average Power, W) (7.81)

Q = VRMS IRMS sin(θv − θi ) (Reactive Power, VAR) (7.82)

7.9.1 Average VS Reactive Power


• Average Power: Useful (actual) power, used by the load.

• Reactive Power: Measure of energy exchange between source and reactive (capacitor OR inductor)
load =⇒ Capacitors/Inductor do not dissipate nor supply power.

7.9.2 Leading VS Lagging From The Sign Of Reactive Power


Because of the nature of the sine function, a leading or lagging power will lead to a change in sign for the
reactive power. Observe (7.82):

1. If Q < 0 =⇒ Capacitive Load (Leading pf, θv < θi )

2. If Q = 0 =⇒ Resistive Load (pf = 1, θv = θi )

3. If Q > 0 =⇒ Inductive Load (Lagging pf, θv > θi )


7.10 Conservation Of AC Power
Like DC circuits, the power supplied by the source should equal to the sum of the power dissipated
by the elements in the circuit. We should expect the same for an AC circuit. Observe the two circuit
configurations:

7.10.1 Conservation Of AC Power In Parallel Configuration


For a parallel circuit, like (a), we note that the phasor current supplied by the source is equal sum of
phasor currents that pass through each impedance:

I = I1 + I2 (7.83)
Hence, the power supplied by the source must equal to the sum of the powers dissipated by each element.
We prove this with the following:

1
S = VI∗ (7.84)
2
1
= V(I∗1 + I∗2 ) (7.85)
2
1 1
= VI∗1 + VI∗2 (7.86)
2 2
= S1 + S2 (7.87)

7.10.2 Conservation Of AC Power In Series Configuration


For a series circuit, like (b), the voltage supplied by the source is equal to the sum of the voltage drops
across each impedance:

V = V1 + V2 (7.88)
We begin with:

1
S = VI∗ (7.89)
2
1
= (V1 + V2 )I∗ (7.90)
2
1 1
= V1 I∗ + V2 I∗ (7.91)
2 2
= S1 + S2 (7.92)
7.11 Power Factor Correction
Sometimes, we may wish to increase the power factor so that, for instance, it is closer to unity. This is
called power correction:

The process of increasing the power factor without altering the voltage and current (and hence,
power) to the original load is called power correction

This power correction is achieved by adding a capacitor in parallel to the original load. In the following
example, the original load was an inductive load :

The addition of a capacitor has decreased the phase difference, θ1 to θ2 (i.e. current lags more because of
this). The impact of this is that as the power factor approaches unity, the reactive power decreases. This
is shown in the following diagram as S1 −→ S2 and as Q1 −→ Q2

The difference in the reactive power is due to the reactive power of the capacitor. From here, we might
want to find what capacitance causes this shift. Suppose if S1 were the apparent power of the original
load:

P = S1 cos(θ1 ) (θ1 = θv − θi ) (7.93)


P
Q1 = S1 sin(θ1 ) = sinθ1 = P tanθ1 (7.94)
cosθ1
We would like increase the power factor from cosθ1 to cosθ2 without changing P. This means that:

Q2 = P tanθ2 (7.95)
This difference must be due to the reactive power of the capacitor itself. Hence:

Qc = Q1 − Q2 = P (tanθ1 − tanθ2 ) (7.96)


However, we also know that in general that Q = IRM S X 2 , for a capacitor, this would mean:

2
Qc = IRM S Xc (7.97)
2
VRM S
= (7.98)
Xc
2
= VRM S Cω (7.99)

Substituting (7.99) into (7.96), and solving for the capacitance, C, we get:

P (tanθ1 − tanθ2 )
C= 2 (7.100)
ωVRM S
Chapter 8

Magnetically Coupled Circuits

Magnetically coupled circuits are circuits which are connected via electromagnetic induction. This is
shown below:

Faraday’s Law states that the voltage induced in an conducting coil, via electromagnetic induction, is due
to the rate of change in flux and the number of turns the coil has:


v=N (8.1)
dt
via the Chain Rule, we can expand (8.1):

dφ di
v=N (8.2)
di dt
di
=L (8.3)
dt

Where L = N dφ
di is defined to be the inductance of the coil in Henries (H)

Circuits are magnetically coupled due to mututal inductance. We shall derive mathematical expressions
for mutual inductance using the scenario above:

64
8.1 Mutual Inductance
Suppose coil (inductor) 1 as N1 turns while coil 2 has N2 turns. Let coil 1 be connected to a current
source, i1 (t).

Voltage Across Coil 1:

dφ1
v1 = N1 (8.4)
dt
dφ1 di1
= N1 (8.5)
di1 dt
di1
= L1 (8.6)
dt

Voltage Across Coil 2:


This is where mutual inductance comes into play. Let the mutual inductance of coil 2 with respect to coil
1 be denoted as M21 . Then, at coil 2:

dφ12
v2 = N 2 (8.7)
dt
dφ12 di1
= N2 (8.8)
di1 dt
di1
= M21 (8.9)
dt
On the other hand, if we swap the current source so that now its on the right-hand loop and supplies the
current i2 (t), we would get:

di2
v2 = L2 (8.10)
dt
di2
v1 = M12 (8.11)
dt

8.1.1 Mutual Inductance Relationship


As long as the inductors are in close proximity and the circuits are driven by time-varying sources, we
can say that:

M12 = M21 = M (8.12)


8.2 Mutual Inductance: Sign Convention

If a current enters the dotted terminal of one coil, the reference polarity of the mutual voltage in
the second coil is positive at the dotted terminal of the second coil

Alternatively:

If a current leaves the dotted terminal of one coil, the reference polarity of the mutual voltage in
the second coil is negative at the dotted terminal of the second coil.

8.3 Mutual Inductance: Circuit Analysis


When analysing circuits, we need to take into account mutual inductance, M , when performing operations
like KVL. Fortunately, we can convert such circuits into easier ones to solve. The procedure is shown
below with an example circuit:
Procedure
1. Observe loop with an independent source and its polarity. Using its polarity, indicate the direction
of the current.

2. Place a dependent source, anywhere next to the first coil, with a polarity arranged to follow direction
of current.

3. The dependent source is a controlled voltage source with a voltage equal to the product of the
impedance associated with the mutual inductance (i.e. j3Ω) and the mesh current in the second
loop (i.e. I2 ). This is shown below:

4. Next, observe the second loop without an input source. Observe the dot notation on the coil in
this loop. The dot indicates the positive terminal of the coil.

5. Place a dependent source anywhere next to the inductor with the same direction of polarity as the
inductor.

6. The dependent source has a voltage equal to the product of the impedance associated with the
mutual inductance (i.e. j3Ω) and the current in the other loop (i.e. I1 ). This is shown below:

7. You can then perform KVL to find the mesh currents and do whatever you need to do with the
solve currents.
8.4 Mutual Inductance: General Circuit Analysis
Unlike the circuit prior which was a magnetically coupled circuit, the following circuit is an example of a
general circuit that exhibits mutual inductance without separation between the loops.

Here, we are conflicted with a 6jΩ inductor that intersects two loops which may make things a bit more
confusing. However, these following simple steps will allow the circuit to be solvable.

1. Add dependent voltage sources next to the inductors. The positive terminal of the dependent voltage
source needs to be on the same side as the dot (i.e. Dot → Left, ∴ +Ve terminal → Left)

2. The impedance associated with the mutual inductance is j3Ω. To determine the current from each
of the dependent sources, observe the following diagram:
(a) Top-Left Inductor: Its voltage must come from the mesh current through the 2nd inductor
coil at the mesh intersection. Using the pre-defined currents for each loop, the intersection
current is given by:

Intersection Current = I2 − I1 (8.13)

Hence the voltage of the first dependent source is:

v1 = 3j(I2 − I1 ) (8.14)

(b) Middle Inductor: The current from the 2nd inductor coil at the mesh intersection must
come from the mesh current in the first loop, I1 . Hence:

v2 = 3jI1 (8.15)

3. Once everything has been labelled and defined, KVL can be performed to solve for the mesh currents.

NOTE: The polarity of the dependent source does not determine the mesh current direction in
each loop. The dependent sources are just a convenient way to include the impedance associated with
the mutual inductance in the circuit analysis.

For simplicity, simply define the mesh current to always be clockwise and use the dot convention from
there afterwards.

8.5 Mutual Inductance: Sum Of Inductances


With the dot convention in mind and also the fact that nearby inductors have a mutual inductance, M ,
we simply cannot add the inductances for the total inductance. Depending on the arrangement, we have
the following:

8.5.1 Series-Aiding Connection

L = L1 + L2 + 2M (8.16)
8.5.2 Series-Opposing Connection

L = L1 + L2 − 2M (8.17)
8.6 Energy In A Coupled Circuit
The energy within an inductor was defined earlier to be:
1
w = Li2 (8.18)
2
In the context of a couple circuit, mutual inductance needs to be included in the total stored energy.
Consider the following circuit where both current enter the dots shown:

In this case, the total energy stored within the coupled circuit is:

1 1
w = L1 i21 + L2 i22 + M i1 i2 (8.19)
2 2

However, if one of the currents enters a dot and the other current leaves a dot, then the total energy
stored by the couple circuit is:

In this case, the total energy stored within the coupled circuit is:

1 1
w = L1 i21 + L2 i22 − M i1 i2 (8.20)
2 2
8.7 Coupling Coefficient

The Coupling Coefficient is given by:


M
k=√ (8.21)
L1 L2
It is a measure of the magnetic coupling between the two coils, where 0 ≤ k ≤ 1

8.7.1 Example Coupling Coefficient Question


Consider the circuit below. Determine the coupling coefficient. Calculate the energy stored in the coupled
inductors at time t = 1s if v = 60cos(4t + 30◦ ) V.

KVL in Loop 1:

10I1 + (10j)I2 + (20j)I1 = 60∠30◦ (8.22)



(1 + 2j)I1 + (j)I2 = 6∠30 (8.23)

KVL in Loop 2:

(16j)I2 − (4j)I2 + (10j)I1 = 0 (8.24)


−6
I1 = I2 (8.25)
5
Subbing (8.23) → (8.21) and solving for I2 :

6
(1 + 2j)(− )I2 + (j)I2 = 6∠30◦ (8.26)
5
I2 = 3.25∠160.60◦ (8.27)

Subbing (8.25) → (8.23) and solving for I1 :

I1 = 3.90∠−19.40◦ (8.28)
Converting from the frequency-domain into the time-domain:

i1 (t) = 3.90cos(4t − 19.40◦ ) (8.29)



i2 (t) = 3.25cos(4t + 160.60 ) (8.30)

Now, we know that ω = 4rad/s. We need to convert this into degrees when t = 1s

720 ◦
 
rad 180
4t = 4 (1s) = 4rad = 4 = (8.31)
s π π
Hence, we can say that at t = 1:

720 ◦
i1 = 3.90cos( − 19.40◦ ) = −3.38 (8.32)
π
720 ◦
i2 = 3.25cos( + 160.60◦ ) = 2.82 (8.33)
π
Since both mesh currents pass into the dot, we use the following equation for the total energy in the
coupled circuit:

1 1
w = L1 i21 + L2 i22 + M i1 i2 (8.34)
2 2
1 1
= (5)(−3.38)2 + (4)(2.82)2 + (2.5)(−3.38)(2.82) (8.35)
2 2
= 20.73J (8.36)
8.8 Transformers

Transformers are a four-terminal device with two (or more) magnetically coupled coils.

The circuit design for a transformer is shown below:

A transformer is said to be linear if the coils are wrapped around a linear material - e.g. air,
plastic, wood etc. that allow for a constant magnetically permeability.

8.9 Input Impedance, Zin & Reflected Impedance, ZR


From the above circuit, we may be interested in trying to find the input impedance, Zin since that governs
the behaviour of the circuit. To do this, we first convert all values to impedances, add dependant sources
and perform KVL in each loop:

KVL Loop 1:

R1 I1 − (jωM )I2 + (jωL1 )I1 = V (8.37)


(R1 + jωL1 )I1 − (jωM )I2 = V (8.38)

KVL Loop 2:

R2 I2 + ZL I2 + (jωL2 )I2 = jωM I1 (8.39)


jωM
I2 = · I1 (8.40)
R2 jωL2 + ZL

Subbing (8.38) −→ (8.36) and solving for


(ωM )2
Zin = (R1 + jωL1 ) + (8.41)
R2 + jωL2 + ZL
The fraction in (8.39) is called the ZR or the Reflected Impedance:

Reflected Impedance:
(ωM )2
ZR = (8.42)
R2 + jωL2 + ZL
General Case: The denominator is simply the sum of impedances in the secondary circuit. It
need not have inductors or capacitors and vice versa.

8.10 Ideal Transformer

The Ideal Transformer is one which has:

• Coupling coefficient of k = 1

• Infinite self-inductances of the primary and secondary coil.

The details we need to know concerning the ideal transformer, as shown below, are:

• The voltage across each inducting coil

• The turns ratio of the secondary coil to the primary coil

• The ratio between the primary current to the secondary current.

The turns ratio is usually denoted by n and related the voltages, currents and turns the following way:

V2 I1 N2
n= = = (8.43)
V1 I2 N1

Due to the dot convention, the ratios need not be necessarily be positive. The following are simple steps
to determine the sign of the ratio:

1. If V1 and V2 are both positive or both negative at the dotted terminals, use n. Otherwise, use -n.

2. If I1 and I2 both enter into or both leave the dotted terminals, use -n in Eq. (13.55). Otherwise,
use n.
8.11 Eliminating Transformer By Reflection
Circuit analysis is made easier by reflecting the elements on one coil to the other coil. Consider the
following ideal transformer:

For simplicity, we are going to only reflect the secondary coil onto the primary coil. To do this we:

1. Determine the turn ratio n

2. Divide the secondary impedance(s) by n2

3. Divide the secondary voltage by n

4. Multiply the secondary current, I2 by n

Once the reflection has been complete, the final circuit, looks a little something like this:

From here, the usual circuit analysis techniques can be used to determine quantities such as circuit cur-
rent, voltage across and element etc.

NOTE: If the transformer has an external connection between the primary and secondary coils, then the
reflection cannot be applied. The usual KCL and KVL are the only techniques that can be used.
8.12 Step-Up/Step-Down Transformers
By determine the turn ratio we can deduce whether the ideal transformer is a step-up or step-down trans-
former:

A step-up transformer is one which produces a secondary voltage greater than the primary voltage

A step-down transformer is one which produces a secondary voltage less than the primary voltage

8.13 Autotransformer

An autotransformer is a transformer with which both the primary AND secondary coils are in
one winding.

Step-Down Autotransformer

V1 I2 N1 + N2
n= = = (8.44)
V2 I1 N2

Step-Up Autotransformer

V1 I2 N1
n= = = (8.45)
V2 I1 N1 + N2
Chapter 9

Frequency Response, Transfer Function,


Resonant Circuits

So far, AC circuit analysis has been performed such that frequency has remained constant. However, a
change in frequency will give rise to a frequency response in the circuit. This chapter will explore this
response as well as the concept of a transfer function.

9.1 Frequency Response

The frequency response of a circuit is the variation in its behaviour due changes in the frequency
of an input signal.

The frequency response can be expressed a function of frequency via the transfer function

9.2 Transfer Function

The transfer function is a function of frequency. It is expressed as the frequency-dependent ratio


of the phasor output (e.g. voltage/current of an element) to the phasor input (e.g. voltage/current
source).
Y(ω) Output Signal
H(ω) = = (9.1)
X(ω) Input Signal

The transfer function operates on an input signal via a linear network. An example is the op-amp.
Where the transfer function is the absolute value of the gain.

A simple model to illustrate how the transfer function operates on an input signal is shown below:

78
9.3 Examples Of Transfer Functions

9.4 Transfer Function: Example Question


For the RC circuit in Fig. 14.2(a), obtain the transfer function Vo /Vi and its frequency response. Let
vs = Vm cos(ωt).

Solution
We begin by deriving the transfer function for this particular circuit configuration:

Vo
H(ω) = (9.2)
Vs
1
Vs ( R+jωC1 )
jωC
= (9.3)
Vs
1
= (9.4)
1 + jωRC
Next, we want to investigate the frequency response. For simplicity, let s = ωRC, for the sake of
eliminating tedious algebra. We want to convert H into rectangular form:
1
H(ω) = (9.5)
1 + js
1 1 − js
= · (9.6)
1 + js 1 − js
1 − js
= (9.7)
1 + s2 
  
1 s
= −j (9.8)
1 + s2 1 + s2

Taking the modulus of H will allow us to determine how the magnitude of the H changes with respect
to frequency.

   
1 s
|H(ω)| = 2
−j 2
(9.9)
1+s 1+s
s 2  2
1 s
= + (9.10)
1 + s2 1 + s2
1
=√ (9.11)
1 + s2
1
=√ (9.12)
1 + R2 ω 2 C 2
1
=r (9.13)
2
1 + ω1
R2 C 2

1 1
=q (ωo = ) (9.14)
1+ ω2 RC
ωo2

We can also take the argument of H to see how the change in frequency will impact the phase of the
signal:

1 + s2
 
−s
−1
Arg[H(ω)] = tan · (9.15)
1 + s2 1
 
ω
φ = tan−1 (9.16)
ωo

If we plot the |H| and φ with respect to frequency, we obtain the following graphs (otherwise known as
bode plots):
9.5 Resonance

In an RLC circuit, resonance occurs when the capacitive and inductive reactance are equal. At
that point, the impedance becomes purely resistive.

The resonant behaviour of an RLC circuit depends on whether it is in a series or parallel configuration.
In this section, we will explore both.

9.6 Series RLC Resonance


Observe the following circuit in the frequency domain with all elements in series with an input voltage
source:

9.6.1 Obtaining The Resonant Frequency, ωo


To obtain ωo , we need to solve for the total impedance of the circuit. We can do this by first performing
KVL:

1
RI + (jωL)I + ( )I = Vs (9.17)
jωC
 
j
I R + jωL − = Vs (9.18)
ωC
V
The total impedance is obtained when Z = I. Hence:
 
Vs 1
Z= = R + j ωL − (9.19)
I ωC

Recall that the circuit exhibits resonance when the total impedance is purely resisitive. Hence, the imagi-
nary component of the total impedance must be equal to 0. The frequency corresponding to this condition
is called the resonant frequency:
1
ωo L − =0 (9.20)
ωo C
Hence, solving for ωo :

1
ωo = √ (9.21)
LC
OR as frequency in Hertz (Hz), the resonant frequency can be expressed as:
1
fo = √ (9.22)
2π LC

9.6.2 Maximum & Half-Power Frequencies


In the context of phasors, the average power is given by: P (ω) = 12 I 2 R. However, the maximum average
power occurs when I = VRm . Hence:

Vm2
P (ω) = (9.23)
2R
Of more importance are the half-power frequencies. These frequencies are responsible for the band-
width of a signal which will be discussed later. These frequencies occur at half the maximum average
power.

Let the half-power frequencies be ω1 and ω2 . Hence:

Vm2
P (ω1 ) = P (ω2 ) = (9.24)
4R
These half-power frequencies are related to the capacitance, inductance and resistance through the fol-
lowing equation:

The half-power frequencies are given as:


s 2
R R 1
ω1,2 =∓ + + (9.25)
2L 2L LC
9.6.3 Bandwidth

The bandwidth, B, of a signal is the difference in the half-power frequencies:

B = ω2 − ω1 (9.26)

The bandwidth is also related to the elements in an RLC circuit with the following equation:
R
B= (9.27)
L

9.6.4 Quality Factor

The Quality Factor, Q, of a signal is the ratio between the resonant frequency and the bandwidth:
ωo ωo L 1
Q= = = (9.28)
B R ωo RC
However, it is also the ratio of the peak energy stored in the circuit to the energy dissipated by the
circuit in one period of the resonance.
Peak Energy In Circuit
Q = 2π · (9.29)
Energy Dissipated By Circuit In One Period Of Resonance

The quality factor is related to the resistance, capacitance, inductance and bandwidth with the following
expression:

R
B= (For Series) (9.30)
L
ωo
= (9.31)
Q
= ωo2 RC (9.32)
9.7 Parallel RLC Resonance
The circuit configuration is shown below:

Most chracteristics of the series circuit defined prior are much the same for a parallel with a few minor
differences:

Half-Power Frequencies:
s 2
1 1 1
ω1,2 =∓ + + (9.33)
2RC 2RC RC

Bandwidth
1
B = ω2 − ω1 = (9.34)
RC
Quality Factor
ωo R
Q= = = ωo RC (9.35)
B ωo L

9.8 Half-Power Frequencies Independent Of Circuit Configuration


The expression for half-power frequencies can differ depending on if the circuit is in a series or parallel
configuration. Expressing the frequencies in terms of the quality factor allow both configurations to have
the same expressions for the half-power frequencies:

s  2
ωo 1
ω1,2 = ∓ωo · + 1+ (9.36)
2Q 2Q

9.9 Summary Of Important Equations For Resonant RLC Circuits


The equations have been summarised in the table on the next page. Notice some similarities between the
series and parallel analogue that may make the memorisation easier.
9.10 Filters

Filters are circuits that are designed to allow certain frequencies to pass through and re-
ject/attenuate others.

There are two types of filters to consider:

• Passive: Filter circuits that only contain R, L and C.

• Active: Filter circuits that have R, L and C as well as op-amps and transistors.

No matter is the filter if the filter is considered passive or active, there are 4 main types of filters to
consider:

1. Low-Pass Filter: Permits low-frequency signals and rejects frequencies after the corner/cutoff
frequency, ωc

2. High-Pass Filter: Permits high-frequency signals and rejects frequencies prior ωc

3. Bandpass Filter: Permits frequencies within a certain frequency band (ω1 → ω2 ) and rejects/attenuates
others.

4. Bandstop Filter: Permits frequencies outside of the frequency band and rejects frequencies within
the band.

The impact of these filters on the transfer function, H as a function of frequency, are shown in the graphs
below, respectively, from (a) to (d):
9.11 Passive Filters
These are the first set of filters to discuss. All elements have been converted to phasors and our job is to
find the transfer function and hence the behaviour when ”|H(ω)| VS ω”.

9.12 Low-Pass Filter

Low-Pass Filter: Permits low-frequency signals and rejects frequencies after the corner/cutoff
frequency, ωc . It allows frequencies from dc up till the cut off-frequency.

Once converting everything to phasor form, we can proceed to find the transfer function with the output
voltage signal taken from the capacitor:

 1

jωC
Vs 1
R+ jωC
Vo
H(ω) = = (9.37)
Vi Vs
1
= (9.38)
1 + jRωC
The modulus of the transfer function is given by:
   
1 RωC
|H(ω)| = −j (9.39)
1 + (RωC)2 1 + (RωC)2
Setting the the modulus equal to √1 allows us to solve for the cutoff/corner frequency, ωc :
2

1
ωc = (9.40)
RC
9.13 High-Pass Filter

High-Pass Filter: Permits high-frequency signals and rejects frequencies before the corner/cutoff
frequency, ωc

Designed to pass all frequencies that are above the cutoff/corner frequency, ωc

Once all circuit elements have been converted to phasor form, we can proceed to find the transfer function.
The output voltage is taken from the resistor as shown in the circuit. For convenience:

R
H(ω) = j
(9.41)
R− ωC

Using the same process as for the low-pass filter, the cutoff frequency can be found to be:

1
ωc = (9.42)
RC

Again, this occurs when the modulus of the transfer function is equal to √1 . The graph of the ideal VS
2
actual behaviour of the high-pass filter is shown below:
9.14 Bandpass Filter

Bandpass Filter: Permits frequencies within a certain frequency band for some ω such that
ω1 < ωo < ω2

ωo is called the centre frequency.

We now find the transfer function such that the output voltage is taken with respect to the resistor. In
the frequency-domain, we get:

R
H(ω) =   (9.43)
1
R + j ωL − ωC

We then take the modulus of the transfer function and set it to √1 . We then solve for the centre frequency,
2
ωo which is:

1
ωo = √ (9.44)
LC
9.15 Bandstop Filter

Bandstop Filter: Eliminates frequencies within a frequency band for some ω such that
ω1 < ωo < ω2

ωo is called the centre frequency.

We now find the transfer function such that the output voltage is taken with respect to the inductive
AND capacitive (LC) load. Doing this, we get:

 
1
j ωL − ωC
H(ω) = 1 (9.45)
R + j(ωL − ωC )

Taking the modulus of the transfer function and solving for ωo , we find the centre frequency of the fre-
quency band to be:

1
ωo = √ (9.46)
LC
9.16 Identifying The Type Of Filter
To identify the behaviour of the filter and, hence, the type of filter it is, we need to investigate what
happens to H(ω) as:

• ωo −→ 0

• ωo −→ ∞

The behaviour of each filter for these two cases is shown in the table below:

9.17 General Case For Transfer Function At Corner/Centre Frequency


The table above shows that at the corner/centre frequency, the transfer function should equate to √1 .
2
However, in general, the corner frequency occurs when:
1
H(ωc ) = √ H(∞) (9.47)
2
Or in terms of the magnitude of the transfer function:
1
|H(ωc )| = √ |H(∞)| (9.48)
2
The tabulated results are valid if and only if H(∞) = 1. This the ideal case for the ideal filter.
9.18 Active Filters
Active Filters, perform the same functions as the passive filters discussed previously. However,
Chapter 10

Introduction To Laplace Transform,


Applications And State Variables

93
Chapter 11

Fourier Series & Fourier Transform

94
Chapter 12

Two-Port Networks

95

Вам также может понравиться