You are on page 1of 46

Will Moore

MT 11 & MT12

CIRCUIT ANALYSIS II
(AC Circuits)

Syllabus
Complex impedance, power factor, frequency response of AC networks
including Bode diagrams, second-order and resonant circuits, damping
and Q factors. Laplace transform methods for transient circuit analysis
with zero initial conditions. Impulse and step responses of second-order
networks and resonant circuits. Phasors, mutual inductance and ideal
transformers.

Learning Outcomes
At the end of this course students should:
1. Appreciate the significance and utility of Kirchhoff’s laws.
2. Be familiar with current/voltage relationships for resistors, capacitors
and inductors.
3. Appreciate the significance of phasor methods in the analysis of AC
circuits.
4. Be familiar with use of phasors in node-voltage and loop analysis of
circuits.
5. Be familiar with the use of phasors in deriving Thévenin and Norton
equivalent circuits
6. Be familiar with power dissipation and energy storage in circuit
elements.
7. Be familiar with methods of describing the frequency response of
AC circuits and in particular
8. Be familiar with the Argand diagram and Bode diagram methods
9. Be familiar with resonance phenomena in electrical circuits
10. Appreciate the significance of the Q factor and damping factor.
11. Appreciate the significance of the Q factor in terms of energy
storage and energy dissipation.
12. Appreciate the significance of magnetic coupling and mutual
inductance.
13. Appreciate the transformer as a means to transform voltage, current
and impedance.
14. Appreciate the importance of transient response of electrical
circuits.
15. Be familiar with first order systems
16. Be familiar with the use of Laplace transforms in the analysis of the
transient response of electrical networks.
17. Appreciate the similarity between the use of Laplace transform and
phasor techniques in circuit analysis.

2
Circuit Analysis II WRM MT11

AC Circuits

1. Basic Ideas

Our development of the principles of circuit analysis in Circuit Analysis I


was in terms of DC circuits in which the currents and voltages were
constant and so did not vary with time. We will now extend this analysis
to consider time varying currents and voltages. In our initial discussions
we will limit ourselves to sinusoidal functions. We choose this special
case because, as you have now learnt in P1, it allows us to make use of
some very powerful and helpful mathematical techniques. It is a
common waveform in nature and it is easy to generate in the lab.
However as you have also learnt in P1, any waveform can be expressed
as a weighted superposition of sinusoids of different frequencies and
hence if we analyse a linear circuit for sinusoidal functions we can, by
appropriate superposition, handle any function of time.

Let's begin by considering a sinusoidal variation in voltage

v = Vm cos ωt

3
in which ω is the angular frequency and is measured in radians/second.
Since the angle ωt must change by 2π radians in the course of one
period, T, it follows that

ωT = 2π

1
However the time period T = where f is the frequency measured in
f
Hertz. Thus


ω= = 2πf
T

This is a simple and very important relationship. We naturally measure


frequency in Hz – the mains frequency in the UK is 50Hz – and it is easy
to measure the time period, T (= 1 f ) from an oscilloscope screen.
However as we will soon see, it is mathematically more convenient to
work in terms of the angular frequency ω. Mistakes may be easily made
because in practice the word frequency is commonly used to refer to
both ω and f. It is important in calculations to make sure that if ω
appears, then the correct value for f = 50 Hz, say, is ω = 100π rads/sec.
A simple point to labour I admit, but if I had a pound for every time
someone forgets and substitutes ω = 50 . . . . . . . . !!

4
Circuit Analysis II WRM MT11

In our example above, v = Vm cos ωt , it was convenient that


v = Vm at t = 0 . In general this will not be the case and the waveform will
have an arbitrary relationship to the origin t = 0 or, equivalently the origin
may have been chosen arbitrarily and the voltage, say, may be written in
terms of a phase angle, φ, as

v =V m cos (ωt − φ )

Alternatively, in terms of a different phase angle, ψ, the same waveform


can be written

v =V m sin(ωt +ψ )

where

ψ =π 2 −φ

5
The phase difference between two sinusoids is almost always measured
in angle rather than time and of course one cycle (i.e. one period)
corresponds to 2π or 360°. Thus we might say that the waveform above
is out of phase with the earlier sinusoid by φ. When φ = ± π 2 we say
that the two sinusoids are said to be in quadrature. When φ = π the
sinusoids are in opposite phase or in antiphase.

2. RMS Values

We refer to the maximum value of the sinusoid, Vm, as the “peak” value.
On the other hand, if we are looking at the waveform on an oscilloscope,
it is usually easier to measure the “peak-to-peak” value 2Vm, i.e. from the
bottom to the top. However, you will notice that most meters are
calibrated to measure the root-mean-square or rms value. This is found,
as the name suggests, for a particular function, f, by squaring the
function, averaging over a period and taking the (positive) square root of
the average. Thus the rms value of any function f(x), over the interval x
to x+X, where X denotes the period is

1 x+X 2
frms =
X x
∫ f (y )dy

For our sinusoidal function v = Vm cos ωt

6
Circuit Analysis II WRM MT11

The average of the square is given by

1T 2
∫ Vm cos 2 ωt dt
T0

where the time period T = 2π ω . At this point it's probably easiest to


1
change variables to θ = ωt and to write cos2 θ = (1+ cos 2θ). Thus the
2
mean square value becomes

1 Vm2 2π Vm2
2 2π 0
∫ (1+ cos 2θ)dθ = 2

The root mean square value, which is simply the positive square root of
this, may be written as
Vrms = Vm /√2 ≈ 0.7 Vm.

Since we nearly always use rms values in our AC analysis, we assume


rms quantities unless told otherwise so by convention we just call it V as
in:

7
V = Vm /√2.

So, for example, when we say that the UK mains voltage is 230V what
we are really saying is that the rms value 230V. Its peak or maximum
value is actually 230√2 ≈ 325 V.

To see the real importance of the rms value let's calculate the power
dissipated in a resistor.

Here the current is given by i = v R

Vm
i= cos ωt = Im cos ωt
R

where I m =Vm R and the power, p = vi , is given by

Vm2
p= cos2 ωt
R

8
Circuit Analysis II WRM MT11

If we want to calculate the average power dissipated over a cycle we


must integrate from ωt = 0 to ωt = ωT = 2π . If we again introduce θ = ωt ,
the average power dissipated, P, is given by

1 Vm2 2π
P= . ∫ cos2 θ dθ
2π R 0

1 Vm2 1 2π
P=
2π R
.
20
∫ (1+ cos 2θ)dθ

P =Vm2 2R

If we now introduce the rms value of the voltage V =Vm 2 then the
average power dissipated may be written as

P =V 2 R

Indeed if the rms value of the current I = Im 2 is also introduced then

P =V 2 R = I 2R

which is exactly the same form of expression we derived for the DC


case.

Therefore if we use rms values we can use the same formula for the
average power dissipation irrespective of whether the signals are AC or
DC.

9
10
Circuit Analysis II WRM MT11

3. Circuit analysis with sinusoids

Let us begin by considering the following circuit and try to find an


expression for the current, i, after the switch is closed.

The Kirchhoff voltage law permits us to write

di
L + Ri = Vm cos ωt
dt

This is a linear differential equation, which you know how to solve.


We begin by finding the complementary function, from the homogeneous
equation:

di
L + Ri = 0
dt
which yields the solution:

i = A exp − (Rt L )

11
We now need to find the particular integral which, for the sinusoidal
"forcing function" Vm cos ωt , will take the form B cos ωt + C sin ωt . Thus
the full solution is given by

i (t ) = A exp − (Rt L ) + B cos ωt + C sin ωt

We see that the current consists of a "transient" term, A exp − (Rt L ),


which eventually decays and becomes negligible in comparison with the
"steady state" response. The transient response arises because of the
sudden opening or closing of a switch but we will concentrate here on
the final sinusoidal steady state response. How long do we have to wait
for the steady state? If for example R =100 Ω and L = 25mH then

R L = 4 ×10−3 sec −1 and so after only 1ms exp − Rt L = exp (− 4 )= 0.018


and so any measurements we are likely to make on this circuit will be
truly 'steady state' measurements. Thus our solution of interest reduces
to

12
Circuit Analysis II WRM MT11

i = B cos ωt + C sin ωt

In order to find B and C we need to substitute this expression back into


the governing differential equation to give

ωL{C cos ωt − B sin ωt }+ R{B cos ωt + C sin ωt }= Vm cos ωt

It is now a simple matter to compare coefficients of cosωt and sinωt to


obtain expressions for B and C which lead, after a little algebra, to

Vm
i= 2
{R cos ωt + ωL sin ωt}
R 2 + (ωL )

If we now introduce the inductive reactance X L (= ωL ) we can write this


equation as

Vm ⎧⎪ R XL ⎫⎪
i= ⎨ cos ωt + sin ωt ⎬
R 2 + X L2 ⎪⎩ R 2 + X L2 R 2 + X L2 ⎪⎭

The expression in curly brackets is of the form

cos φ cos ωt + sin φ sin ωt = cos(ωt − φ)

and hence
Vm
i=
2 2
(
cos ωt − ϕ )
R +X L

where

13
"X %
ϕ = tan $ L '
−1

#R &

Thus we see that the effect of the inductor has been to introduce a
phase lag φ between the current flowing in the circuit and the voltage
source. Similarly the ratio of the maximum voltage to the maximum

current is given by R 2 + X L2 which since it is a combination of


resistance and reactance is given the new name of impedance.

It is apparent that we could solve all networks containing combinations


of resistors, inductors and capacitors in this way. We would end up with
a series of simultaneous equations to solve – just as we did when
analysing DC circuits – the problem is that they would be simultaneous
differential equations which, given the effort we went through to solve
one equation in the simple example above, would be very tedious and
therefore rather error-prone. Fortunately there is an easier way.

We are saved because the differential equations we have to solve are


linear and hence the principle of superposition applies. This tells us
that if a forcing function v1(t) produces current i1(t) and a forcing function
v 2 (t ) provides current i 2 (t ) then v 1(t ) + v 2 (t ) produces i1(t ) + i 2 (t ). The
trick then is to choose a more general forcing function v (t ) = v 1(t ) + v 2 (t ) in
which, say, v 1 (t ) corresponds to Vm cos ωt and which made the
differential equation easy to solve. We achieve this with complex
algebra.

You should know that

14
Circuit Analysis II WRM MT11

exp j ωt =cos ωt + j sinωt .


where j = √(-1), [electrical engineers like to use i for current]

so let’s solve the differential equation with the general forcing function

v (t ) =Vm exp j ωt =Vm cos ωt + j Vm sinωt


where

v 1 (t ) =Re v (t ) =Re {Vm exp j ωt } =Vm cos ωt


{ } .

The solution will be of the form


i (t ) = I exp j ωt

where 𝐼 = 𝐼 exp −j∅ will, in general, be a complex number. Then in


order to find that part of the full solution corresponding to the real part of
the forcing function, Vm exp j ωt we merely need to find the real part of

i (t ). Thus

i 1 (t ) =Re {I exp jωt } =Re I exp j (ωt − ϕ )


{ }
(
= I cos ωt − ϕ )
Let's illustrate this by returning to our previous example where we tried
to solve:

15
di
L + Ri = Vm cos ωt
dt

Now, instead, we solve the more general case:

di
L + Ri =Vm exp j ωt
dt

and take the real part of the solution. As suggested above an


appropriate particular integral is i = I exp j ωt which leads to

j ωL I exp j ωt + RI exp j ωt =Vm exp j ωt

The factor exp j ωt is common and hence

(R + j ωL )I =Vm

in which R + j ωL may be regarded as a complex impedance. The


complex current I is now given by

Vm Vm
I= = exp − j φ
R + j ωL 2
R + (ωL )
2

with φ = tan−1 (ωL R ) and hence

Vm
i (t ) = Re{I exp j ωt }= cos(ωt − φ)
2 2
R + (ωL )

16
Circuit Analysis II WRM MT11

which, thankfully, is the same solution as before but arrived at with


considerably greater ease.

Let us be clear about the approach. We have

(i) introduced a complex forcing function Vm exp j ωt knowing that in


reality the voltage source must be real i.e. Re{Vm exp j ωt }.

(ii) We solved the equations working with complex voltages and


complex currents, V exp j ωt and I exp j ωt (or rather V and I since the

time dependence exp jωt cancelled out).

(iii) Since the actual voltage is given by Re{Vm exp j ωt } the actual

current is given by Re{I exp j ωt }= Re{I exp j ψ exp j ωt }= I cos (ωt + ψ ).

(iv) Since the differential of exp j𝜔𝑡 − ∅ = exp j𝜔𝑡 . exp −j∅ is simply
𝑗𝜔. exp j𝜔𝑡 . exp −j∅ and since we always take exp j𝜔𝑡 out as a
common factor, you may see now that our differential equations turn into
polynomial equations in jω (and you knew how to solve these at GCSE!)

This is a very powerful approach that will permit us to solve AC circuit


problems very easily.

17
Example

Let's now do an example to show, formally, how we can solve AC


problems. Let's imagine we want to find the steady state current, i2,
flowing through the capacitor in the following example

The two KVL loop equations may be written

Ldi 1
Ri 1 + + R (i 1 − i 2 ) = E m cos (ωt + α )
dt
and

1
R (i 2 − i 1) + ∫i 2
dt =0
C

Replacing Em cos(ωt + α) by Em exp j (ωt + α)= E1 exp j ωt where


E1 = E m exp jα and further introducing I1 and I2 via
i1 = I1 exp j ωt and i 2 = I 2 exp j ωt we obtain
18
Circuit Analysis II WRM MT11

(2R + j ωL )I1 − R I 2 = E1
⎛ 1 ⎞
⎜⎜ R + ⎟⎟I 2 − R I1 = 0
⎝ j ωC ⎠

and, after a little algebra

E1 E exp j α
I2 = = m
L ⎡ 2 ⎤ M+ jN
R+ + j ⎢ωL −
CR ⎣ ωC ⎥⎦

where M = R + L CR and N = ωL − 2 ωC . We note that this may be


written, introducing tan θ = N M

Em
I2 = exp j (α − θ)
2 2
M +N

and hence the actual current i 2 = Re(I 2 exp j ωt ) may be written as

Em
i 2 (t ) = cos (ωt + α − θ)
2 2
M +N

19
4. Phasors

We have just introduced a very powerful method of circuit analysis. In


essence we have introduced the use of complex quantities to represent
sinusoidal functions of time. The complex number A exp jφ (often written

A∠φ ) when used in this context to represent A cos (ωt + φ) is called a


phasor. Since the phase angle φ must be measured relative to some
reference we may call the phasor A∠0 the reference phasor.

Since the phasor, A exp j φ , is complex it may be represented in


Cartesian form x + jy just like any other complex quantity and thus

x = A cos φ
y = A sin φ
A= x2 + y 2

Further since the phasor is a complex quantity it is very easy to display it


on an Argand diagram (also in this context called a phasor diagram).
Thus the phasor A exp j φ is drawn as a line of length A at an angle φ to
the real axis.

20
Circuit Analysis II WRM MT11

We emphasise that this is a graphical representation of an actual


sinusoid A cos(ωt + φ). The rules for addition, subtraction and
multiplication of phasors are identical to those for complex numbers.

Thus addition:

For multiplication it is easiest to multiply the magnitudes and add the


phases. Consider the effect of multiplying a phasor by j

j A exp j φ = exp j π 2 A exp j φ = A exp j (φ + π 2)

which causes the phasor to be rotated by 90o.

21
Similarly, dividing by j leads to

1
A exp j φ = exp( − j π 2) A exp j φ = A exp j (φ − π 2)
j

i.e. a rotation of –90o.

We finally note that it is usual to use rms values for the magnitude of
phasors.

22
Circuit Analysis II WRM MT11

5. Phasor relations in passive elements

Consider now a voltage Vm cos ωt applied to a capacitor. As we have


indicated we elect to use the complex form Vm exp j ωt and so omit the
"real part" as we calculate the current via

dv d
i =C =C (Vm exp j ωt )= j ωC Vm exp j ωt
dt dt

If we now drop the exp j ωt notation and write the voltage phasor Vm as
V and the current phasor as I we have

1
I = j ωC V or V = I
j ωC

In terms of a phasor diagram, taking the voltage V as the reference

from which we confirm two things we already knew

23
1
(i) the ratio of the voltage to the current is - the reactance.
ωC

(ii) the current leads the voltage by 90°. The pre-multiplying factor j
describes this.

For the inductance an analogous procedure leads to

V = j ωL I

Where the reactance is now jωL and, if we now take, say, the current as
the reference phasor we have

and here the current lags the voltage by 90°.

24
Circuit Analysis II WRM MT11

[It is important to get these relationships the right way around and as a
check we may use the memory aid “CIVIL” – in a capacitor, the current
leads the voltage CIVIL and in an inductor, the current lags the voltage
CIVIL.]

Finally for a resistor we know that the current and voltage are in phase
and hence, in phasor terms

V =I R

25
6. Phasors in circuit analysis

We are now in a position to summarise the method of analysis of AC


circuits.

(i) ( )
We include all reactances as imaginary quantities j ω L = j X L for
an inductor and 1 j ω C (= − j X c ) for a capacitor.

(ii) All voltages and currents are represented by phasors, which usually
have rms magnitude, and one is chosen as a reference with zero phase
angle.

(iii) All calculations are carried out in complex notation.

(iv) The magnitude and phase of, say, the current is obtained as
I exp j φ . This can, if necessary, be converted back into a time varying
expression 2 I cos (ωt + φ).

26
Circuit Analysis II WRM MT11

Suppose we wish to find the current flowing through the inductor in the
circuit below

The reactances have been calculated and marked on the diagram. The
left hand voltage source has been chosen as reference and provides
10V rms. The right hand source produces 5V rms but at a phase angle
of 37° with respect to the 10V source. If we introduce phasor loop
currents I1 and I2 as shown then we may write KVL loop equations as

10 = 5 I1 + j 10 (I1 − I 2 )
5 exp j 37o = 4 + j 3 = − (I 2 − I1 ) j 10 − (− j 5)I 2

where we have noted that 5 exp j 37o = 5 cos 37o + j 5 sin 37o = 4 + j 3 . It
is routine to solve these simultaneous equations to give

(7 − j ) 6.5 + j 8
I1 = and I 2 =
12.5 12.5

and hence the current I = I1 − I 2 becomes

-
27
0.5 j 9
I= = 0.72 ∠ − 86.8o = 0.72 exp − j 86.8o
12.5

Since rms values are involved, if we want to convert this into a function
of time we must multiply by 2 to obtain the peak value. Thus

(
i (t ) = 1.02 cos ωt − 86.8o )
In our example we do not know the value of ω but it was accounted for in
the value of the reactances. Since everything is linear and the sources
are independent it would be a good exercise for you to check this result
by using the principle of superposition.

We have used mesh or loop analysis in our examples so far. It is, of


course equally appropriate to use node-voltage analysis if that looks like
an easier way to solve the problem.

As an example let's suppose we would like to find the voltage V in the


circuit below where the reactances have been calculated corresponding
to the frequency, ω, of the source

28
Circuit Analysis II WRM MT11

It's probably as easy as anything to introduce two phasor node voltages


V1 and V. The two node voltage relationships may be written as

V1 − 10 V1 −V V1 − 0
+ + =0
10 j5 − j5

and

V − V1 V − 0 V −0
+ + =0
j5 5 − j10 5 + j10

We note that in writing these equations no thought was given to whether


currents flowing into or out of the nodes were being considered. As in
the DC case it is merely necessary to be consistent. It is now
straightforward to solve these two simultaneous equations to yield

V1 = 10 ∠ − 71.6o or, if the time domain result is required, remembering

(
that the voltage supply is 10V rms then v1(t ) = 20 cos ωt − 71.6o . )
29
7. Combining impedances

As we have seen before the ratio of the voltage to the current phasors is
in general a complex quantity, Z, which generalises Ohms law, in terms
of phasors, to

V =Z I

where Z in general takes the form

Z = Re + j X e

where the overall effect is equivalent to a resistance, Re, in series with a


reactance Xe. If Xe is positive the effective reactance is inductive
whereas negative values suggest that the effective reactance is
capacitative.

Consider the circuit below

30
Circuit Analysis II WRM MT11

⎛ 1 ⎞
V = ⎜⎜ R + j ω L + ⎟ I
⎝ j ω C ⎟⎠

Thus the combined impedance Z is given by

⎛ 1 ⎞
Z = R + j ⎜ ωL − ⎟ = R + j X
⎝ ωC ⎠

This may be visualised on an Argand or phasor diagram

We note that the reactance may be positive or negative according to the

relative values of ωL and 1 ωC . Indeed at a frequency ω= LC we see


that X = 0 and that the impedance is purely resistive. We will return to
this point later.

31
It is straightforward to show, and hopefully intuitive, that all the DC rules
for combining resistances in series and parallel carry over to
impedances. Thus if we have n elements in series, Z1, Z 2 , Z 3 … Z n

Where
n
Zeff = Z1 + Z2 + Z3 + … Zn = ∑ Zi
i =1

and similarly for parallel elements

1 1 1 1 N 1
= + + +…= ∑
Z eq Z1 Z 2 Z 3 i =1 Z i

We note that the inverse of impedance, Z, is known as admittance, Y.


Thus, as in the DC case it is sometimes more convenient to write

n
Yeq = ∑Yi
i =1

32
Circuit Analysis II WRM MT11

and finally since Y is also a complex number it may be written

Y =G + j B

where G is a conductance and B is known as the susceptance.

33
Example

Find the equivalent impedance of the circuit below

Using the usual combination rules gives

1
(R + j ω L)
Z Z j ωC
Z= 1 2 =
Z1 + Z 2 1
R + j ωL +
j ωC

Which we can simplify to

R + j ωL
Z= 2
1− ω LC + j ω C R

34
Circuit Analysis II WRM MT11

8. Operations on phasors

We have just introduced a method of analysing AC circuits in terms of


complex currents and voltages. This method inevitably involves the
manipulation of complex phasor quantities and so we list below the
results for manipulating these quantities which are, course, simply the
standard rules for complex numbers. Sometimes it is easier to use the
a+jb notation and sometimes the r exp jθ = r∠θ notation is easiest. We
summarise below the important relationships.

Addition and Subtraction

If
I1 = a + jb and I 2 = c + jd
then
I1 ± I 2 = a ± c + j (b ± d )

where the real and imaginary parts add/subtract

35
Multiplication
Here it is easiest by far to use the r ∠θ rotation.

If I1 = r1 ∠θ1 = r1 exp jθ1 and I 2 = r2 exp jθ2

then I1 I 2 = r1 r2 exp j (θ1 + θ2 ) = r1 r2 ∠θ1 + θ2

when we see the amplitudes multiply and the arguments add

For division we have

I1 r1 exp j θ1 r1 r
= = exp j (θ1 − θ 2 ) = 1 ∠θ1 − θ2
I 2 r2 exp j θ2 r2 r2

the amplitudes divide and the arguments subtract.

36
Circuit Analysis II WRM MT11

Complex conjugates also appear.

If
I = a + j b = r exp j θ = r ∠θ

then the complex conjugate I*, is given by

I * = a − j b = r exp − j θ = r ∠ − θ

from which we see

I + I * = 2 Re{}
I ; I − I * = 2 j Im{}
I

where Re{ } denotes the real part the Im { } denotes the imaginary part.

37
Rationalising

We are often confronted with expressions of the form

a+ j b
c+ jd

And sometimes we wish to rationalise them. We do this in one of two


ways. The first is to multiply top and bottom by c − j d . This gives

a+ j b a+ j b c− jd ⎛ a c + bd ⎞ ⎛ bc − ad ⎞
= = ⎜⎜ 2 ⎟ +
2 ⎟
j ⎜⎜ 2 ⎟
2 ⎟
c+ jd c+ jd c− jd ⎝ c + d ⎠ ⎝ c + d ⎠

Alternatively, we can write a + jb as r1 exp j θ1 with r1 = a 2 + b 2 and

tan θ1 = b a . Similarly c + jd may be written as r2 exp j θ 2 and hence

a + jb r1 exp jθ1 r1 r
= = exp j (θ1 − θ 2 ) = 1 ∠θ1 − θ 2
c + jd r2 exp jθ2 r2 r2

where the amplitudes divide and the arguments subtract.

There is no golden rule as to which approach to take – it is determined


by the problem at hand. However, if you do not have a pressing need to
rationalise the expression, we will see some quite good reasons why we
may often prefer to stick with the factorised form and not rationalise at
all.

38
Circuit Analysis II WRM MT11

9. Phasor diagrams

Although direct calculation is easily carried out using phasors it is


sometimes useful to use a phasor (Argand) diagram to show the
relationships between say a voltage and current phasor graphically. In
this way it is easy to see their relative amplitudes and phases and hence
gain quick insight into how the circuit operates. As a simple example
consider the circuit below

Since the current, I, flows through both elements it is sensible to choose


this as the reference phasor. Having made this choice the voltage drop
across the resistor, VR = IR , whereas that across the capacitor,
Vc = I jωC = − j ωC I . The sum of these voltages must equal V. The
phasor diagram is easily drawn as

39
from which it is clear that the voltage lags the current by an angle φ. The

angle φ may be obtained from the diagram as φ = tan 1


(1 ωCR).

Let's consider another example

We could obtain the relationship between I and V by using the


equivalent impedance derived earlier. However we will use a phasor
diagram to show the various currents and voltages which appear across
the various components. Since the voltage, V, is the same across each
arm it is sensible to choose this as the reference phasor.

The relationships are

I 2 = jωC V V = I1 R + jωL I1 and I = I1 + I 2

The two phasor diagrams are

40
Circuit Analysis II WRM MT11

or, combining onto a single diagram

where φ denotes the phase angle between V and I. In the diagram


above I leads V by φ.

41
10. Thévenin and Norton equivalent circuits

The Thévenin and Norton theorems apply equally well in the AC case.
Here we replace any arbitrarily complicated circuit containing resistors,
capacitors, inductors by a circuit whose behaviour, as far as the outside
world is concerned, is entirely equivalent.

The two choices are the Thévenin equivalent

and the Norton equivalent

The methods for determining V, I and Z are identical to those used in the
DC case. In general

(i) Calculate the open circuit voltage, Voc

(ii) Calculate the short circuit current, Isc

42
Circuit Analysis II WRM MT11

From which
Voc
V =Voc , I = Isc and Z=
Isc

We note, of course that in the absence of dependent sources it is often


easier to "set the sources to zero" and simply calculate the terminal
impedance Zab. We emphasise again that when a voltage source is "set
to zero" it is replaced by a short circuit whereas when a current source is
"set to zero", no current flows, and hence the source is replaced by an
open circuit.

Finally we note that the equivalent impedance Z is also frequently


referred to as either

(i) internal impedance


or
(ii) output impedance.

43
Example

Find the Norton and Thévenin equivalents of

Noting that there are no dependent sources it is easiest to calculate Zab


directly with the voltage source replaced by a short circuit. This is easy
since the circuit then reduces to an inductor in parallel with a resistor.
Thus

j 40.20
Z = Z ab = =16 + j 8
20 + j 40

We now need to calculate the open circuit voltage between a and b.


This is easy since the circuit is essentially a voltage divider

20
Voc = 5 exp j 10 
50
20 + j 40
50 exp j 10 
= 
= 22.3 exp − j 53.4 
5 exp j 63.4
= 22.3 ∠ − 53.4  V

Thus the Thévenin equivalent circuit takes the form

44
Circuit Analysis II WRM MT11

In order to find the Norton equivalent we need to find the current flowing
between the terminals a and b when they are shorted together. In this
case the circuit becomes

and

50 exp j 10 50 exp j 10 


I sc = = 
= 1.25 exp − j 80 = 1.25 ∠ − 80
j 40 40 exp j 90

and hence the Norton equivalent becomes

45
We have elected to find the Norton equivalent directly. However it is
equally possible to transform between Thévenin and Norton equivalents
directly as we did in the DC case. It is left as an exercise to confirm that

Hence we could have worked out the Norton current source in our
example directly from the Thévenin equivalent as

Voc 22.36 exp − j 53.04 


I= =
Z 16 + j 8

22.36 exp − j 53.4 


= = 1.25 exp − j 80  = 1.25 ∠ − 80 
17.89 exp j 26.6

which, of course, is the value we previously calculated.

46

You might also like