Вы находитесь на странице: 1из 21

5

Magnetically Coupled Circuits

A magnetically (inductively) coupled inductor circuit consists of more than one coil of conductive
wire wound on the same magnetic core. It presents the basis for the transformers that are used to
increase/decrease AC (alternating current) voltages/currents. The coil on the input source side and that
on the output load side are called the primary and secondary coils, respectively. The AC voltage on
the primary coil causes the flux linkage to be changed continually, which induces the voltage on the
secondary coil. The input voltage on the primary coil and the output voltage on the secondary coil are
proportional to the number of windings of each coil. What is the difference between DC (direct current)
and AC? AC means an electric current whose magnitude and direction change periodically, while DC
means an electric current whose magnitude and direction do not change periodically.
It is not by coincidence but due to historical background that this chapter on the basic principle of the
transformer falls here between the previous chapters on DC analysis and the next chapters on AC
analysis. Late in the nineteenth century, there was a fierce competition, called the ‘War of Currents’, for
leadership in the growing market of power transmission and distribution over North America between the
two giants of electrical service, the Edison General Electric Company, who had established a DC power
service system, and the Westinghouse Corporation, who had developed an AC power distribution system.
According to websites such as References [W-1] and [W-5], Edison actively campaigned for the selection
of the AC electric chair as a new executioner, hoping that AC would be known to be fatally dangerous
and, thereby, consumers would not want to use AC. Edison even provided the AC generators that were
needed for the first working electric chairs, although Westinghouse refused to sell any AC generators
directly to prison authorities. Despite these attempts by Edison, who had developed the world’s first
viable system of centrally generating and distributing electric power, the DC power system was
completely defeated by the AC power system, due to transformers (developed by William Stanley)
and AC generators/motors (invented by Nikola Tesla). With AC motors making large power use efficient
and transformers stepping up/down the voltages of an AC power system for efficient power transmission/
distribution, AC power technology was able to consolidate its superiority over DC power technology.

5.1 Self-Inductance
An inductor consists of a coiled conducting wire wound around a core, as illustrated in Figure 5.1(a). In
the magnetic circuit made of the core, the magnetic reluctance and its reciprocal, called the permeance,
are determined as

l 1 A
R¼ ½A  turns=Wb and P¼ ¼  ½Wb=ðA  turnsÞ
A R l

respectively, where l ½m and A ½m2  are the (mean) length and cross-sectional area of the flux path
through the core, and  is the permeability of the core material. The magnetomotive force (mmf)

Circuit Systems with MATLAB1 and PSpice1 Won Y. Yang and Seung C. Lee
# 2007 John Wiley & Sons (Asia) Pte Ltd. ISBN: 978-0-470-82232-6
224 Chapter 5 Magnetically Coupled Circuits

Figure 5.1 A model for an inductor

generated by the current i [A] flowing through the N-turn coil is N i ½A  turns and produces a magnetic
flux

Ni
¼ ¼ P N i ½Wb ð5:1Þ
R

through the core in the direction determined by Ampere’s right-hand rule (see Figure 5.1(b)). This is
analogous to the electric current i ¼ V=R produced by an electromotive force (emf) V in the electric
circuit having a resistance R.
The flux linkage of the N-turn coil is defined to be the flux times the number of turns as

l ¼ N  ½Wb  turns ¼ P N 2 i ¼ L i ð5:2Þ

where the (self)-inductance of the coil (inductor) is defined to be the constant of proportionality of the
flux linkage to the current as

l
L½H ¼ ½Wb  turns=A ¼ P N 2 ð5:3Þ
i

The unit of inductance is the henry (denoted by H), named in honor of the American physicist Joseph
Henry (1797–1878). The flux ‘linkage’ stems from the fact that the flux  through the core is linked with
the current i through the N-turn coil.
Faraday’s law states that the change in the flux linkage induces a voltage across the conductor (coil)
linked with the flux, which equals the time rate of change of the flux linkage:

dlðtÞ ð5:2Þ dðtÞ diðtÞ


vðtÞ ¼ ¼ N ¼L ð5:4Þ
dt dt dt

In fact, Faraday’s law together with Lenz’s law is described by the following equation:

dlðtÞ
eðtÞ ¼ 
dt

where the negative sign means that the polarity of the induced voltage (emf) is such that it opposes the
change of the flux linkage; i.e. it generates a current (through the external network) to produce a magnetic
flux in the direction opposing the change of the flux linkage.
5.2 Mutual Inductance 225

[Remark 5.1] Ampere’s Right-Hand Rule on the Direction of the Magnetic Flux Produced by the
Current
Ampere’s right-hand rule describes the direction of the magnetic flux produced by a current flowing
through a conductor. If the conductor is grasped with the right hand in such a way that the thumb points in
the direction of the current, your fingers wrapping around the conductor curl in the direction of the
magnetic flux (see (A1) in Figure 5.1(b)). If you curl the fingers of your right hand around a coil in the
direction of the current, the thumb points in the direction of the magnetic flux (see (A2) in Figure 5.1(b)).

5.2 Mutual Inductance


In this section two coils are considered that are placed in proximity to each other or wound around a core
as in Figure 5.2, where they have N1 turns and N2 turns, respectively. The flux i linking each coil i is the
sum or difference of two components, the leakage flux ii produced by the current through the coil itself
and the mutual flux ij produced by the current through the other coil j:
1 ¼ 11  12 ; 2 ¼ 21 þ 22 ð5:5Þ

where

11 ¼ the leakage flux linking coil 1 and produced by the current i1 through coil 1
12 ¼ the mutual flux linking coil 1 and produced by the current i2 through coil 2
21 ¼ the mutual flux linking coil 2 and produced by the current i1 through coil 1
22 ¼ the leakage flux linking coil 2 and produced by the current i2 through coil 2

Thus the flux linkage of each coil can be written as


ð5:2Þ
l1 ¼ N1 1 ¼ N1 ð11  12 Þ ¼ L1 i1  M12 i2 ð5:6aÞ
ð5:2Þ
l2 ¼ N2 2 ¼ N2 ð21 þ 22 Þ ¼ M21 i1 þ L2 i2 ð5:6bÞ

pffiffiffiffiffiffiffiffiffiffi
with the self-inductances L1 and L2 ; the mutual inductance M12 ¼ M21 ¼ M ¼ k L1 L2 ; ð5:7Þ
M
and the coefficient of coupling k ¼ pffiffiffiffiffiffiffiffiffiffi ð0  k  1Þ ð5:8Þ
L1 L2

Figure 5.2 Relative winding directions of magnetically coupled coils


226 Chapter 5 Magnetically Coupled Circuits

where the signs of the mutual inductance terms are plus or minus depending on whether the fluxes produced
by currents through two coils in the reference directions are additive or subtractive. Consequently, the
induced voltages are the sum or difference of a self-induced one and a mutually induced one as

ð5:4Þ dl1 ð5:6aÞ di1 di2


v1 ðtÞ ¼ ¼ L1 M ð5:9aÞ
dt dt dt
ð5:4Þ dl2 ð5:6bÞ di1 di2
v2 ðtÞ ¼ ¼ M þ L2 ð5:9bÞ
dt dt dt

5.3 Relative Polarity of Induced Voltages and Dot Convention


5.3.1 Dot Convention and Sign of Mutual Inductance Terms
The polarity of a mutually induced voltage relative to a self-induced one across a coil coupled
magnetically with another coil depends on the relative winding and current directions of the two coils.
Since it is cumbersome to draw the winding details as depicted in Figure 5.2, the dot convention is used to
indicate the relative coil winding direction in the following way:

Dot (Marking) Convention:


The self-induced voltage and the mutually induced one are additive, i.e. have the same polarity if
both coil currents enter/leave the dotted or undotted ends of the coils (Figure 5.3(a)). They are
subtractive, i.e. opposed to each other, if one coil current enters/leaves the dotted end of a coil
while the other coil current enters/leaves the undotted end of the other coil (Figure 5.3(b)).

5.3.2 Measurement of the Relative Winding Direction


Figure 5.4(a) shows a testing circuit to determine the relative winding direction of a pair of magnetically
coupled coils that can be indicated by the dot marks. Let the switch be closed so that i1 > 0 through coil 1
(with the dot on the upper side) produces some flux 11 through the core. Then a current i2 is supposed to
be induced through coil 2 in such a direction that it will produce a flux 22 opposing 11 . There are two
possible cases where a dot should be marked on the coil, on its upper or lower side:
1. If the voltmeter indicates a positive secondary voltage v2 > 0, this implies that the current i2 flows
upward through coil 2. Since this current i2 < 0 must have produced the flux 22 opposing 11 (Lenz’s
law), the dot on coil 2 should be marked on the upper side so that i2 < 0 enters the undotted terminal of
coil 2, while i1 > 0 enters the dotted terminal of coil 1 (see Figure 5.4(b1)).
2. If the voltmeter indicates a negative secondary voltage v2 < 0, it implies that the current i2 flows
downward through coil 2. Since this current i2 > 0 must have produced the flux 22 opposing 11

Figure 5.3 The sign of the mutal inductance terms depending on the current reference directions and relative
winding directions of magnetically coupled coils expressed by the dot convention
5.3 Relative Polarity of Induced Voltages and Dot Convention 227

Figure 5.4 To find the relative winding directions of magnetically coupled coils

(Lenz’s law), the dot on coil 2 should be marked on the lower side so that i2 > 0 enters the undotted
terminal of coil 2, while i1 > 0 enters the dotted terminal of coil 1 (see Figure 5.4(b2)).

5.3.3 Measurement of Mutual Inductance


To find the mutual inductance of a pair of coupled coils, there is a need to find the difference between two
resulting inductances measured for the two connections in Figures 5.5(a) and (b). The overall voltage–
current relationship and the resulting inductance of the circuit connected as in Figure 5.5(a) is
   
di1 di2 di1 di2
vðtÞ ¼ v1 ðtÞ  v2 ðtÞ ¼ L1 M  M þ L2
dt dt dt dt
   
ði1 ¼i; i2 ¼iÞ di di di di di
¼ L1  M  M  L2 ¼ ðL1 þ L2  2MÞ
dt dt dt dt dt
La ¼ L1 þ L2  2M ð5:10aÞ

The overall voltage–current relationship and the resulting inductance of the circuit connected as in
Figure 5.5(b) is
   
di1 di2 di1 di2
vðtÞ ¼ v1 ðtÞ þ v2 ðtÞ ¼ L1 M þ M þ L2
dt dt dt dt
   
ði1 ¼i; i2 ¼iÞ di di di di di
¼ L1  M þ M þ L2 ¼ ðL1 þ L2  2MÞ
dt dt dt dt dt
Lb ¼ L1 þ L2  2M ð5:10bÞ

Figure 5.5 Test to measure a mutual inductance


228 Chapter 5 Magnetically Coupled Circuits

Combining these two equations, 1/4 is multiplied by the difference between these two resulting
inductances to obtain the mutual inductance as

jLa  Lb j
M¼ ð5:11Þ
4

5.3.4 Energy in Magnetically Coupled Coils


Let us find the energy stored in a pair of magnetically coupled coils with the self-inductances L1 and L2
and the mutual inductance M. On the assumption of zero initial conditions, the total power delivered to
the two coils is integrated to find the energy as

pX ðtÞ ¼ v1 ðtÞi1 ðtÞ þ v2 ðtÞi2 ðtÞ


   
di1 ðtÞ di2 ðtÞ di1 ðtÞ di2 ðtÞ
¼ L1 M i1 ðtÞ þ M þ L2 i2 ðtÞ
dt dt dt dt
d d d
¼ L1 i1 ðtÞ i1 ðtÞ  M ½i1 ðtÞ i2 ðtÞ þ L2 i2 ðtÞ i2 ðtÞ
dt dt dt
ðt
wX ðtÞ ¼ pX ðÞd
0
ðt  
d d d
¼ L1 i1 ðÞ i1 ðÞ  M ½i1 ðÞ i2 ðÞ þ L2 i2 ðÞ i2 ðÞ d
0 d d d
ð i1 ðtÞ;i2 ðtÞ
¼ ½L1 i1 d i1  M dði1 i2 Þ þ L2 i2 di2 
0
¼ 12 L1 i21  M i1 i2 þ 12 L2 i22 ð5:12Þ

As a by-product, it can be shown that the (magnetic) coupling coefficient k cannot be greater than unity,
based on this energy equation together with the fact that a pair of coupled coils is a passive element and so
its energy can never be negative; i.e. noting that the above energy equation can be written as
 2  
1 pffiffiffiffiffi M 1 M2 2
wX ðtÞ ¼ L1 i1  pffiffiffiffiffi i2 þ L2  i  0 8 i1 and i2
2 L1 2 L1 2

and this must be nonnegative even for i1 ¼ ðM=L1 Þi2 (making the first squared term zero), the desired
result is obtained:
M2 pffiffiffiffiffiffiffiffiffiffi ð5:7Þ pffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffi
L2   0; M 2  L1 L2 ; M L1 L2 ! k L1 L2  L1 L2 ; 0k1
L1

5.4 Equivalent Models of Magnetically Coupled Coils


It is not straightforward to set up mesh equations or node equations for circuits containing coupled coils.
It would be much easier if a pair of coupled coils is replaced by an equivalent that contains dependent
voltage sources to account for the mutual inductance effect (Figure 5.6.1) or another equivalent that
consists of three uncoupled inductors (Figure 5.6.2). Especially, the equivalents in Figures 5.6.1 and
5.6.2(a) are good for setting up mesh equations and the equivalent in Figure 5.6.2(b) is suitable for setting
up node equations, where the signs of M should be the upper or lower ones of the double signs ( or )
depending on whether the dots of both coils are on the same or opposite sides.
5.4 Equivalent Models of Magnetically Coupled Coils 229

Figure 5.6.1 The equivalent model of two magnetically coupled coils with dependent sources

Figure 5.6.2 The equivalent model of two magnetically coupled coils with no dependent source

Note. Coupled inductors are mainly used for AC applications since coils are just like short-circuits in the DC steady
state (Remark 3.2(2)).

5.4.1 T-Equivalent Circuit


Taking the Laplace transform of the time-domain voltage–current relationship (Equations (5.9)) of a pair
of magnetically coupled coils yields
    
V1 ðsÞ þ L1 i1 ð0Þ  M i2 ð0Þ sL1 sM I1 ðsÞ
¼
V2 ðsÞ þ L2 i2 ð0Þ  M i1 ð0Þ sM sL2 I2 ðsÞ
    
zero initial conditions V1 ðsÞ sL1 sM I1 ðsÞ
! ¼
V2 ðsÞ sM sL2 I2 ðsÞ
  
sðL1  M  MÞ sM I1 ðsÞ
¼ ð5:13Þ
sM sðL2  M  MÞ I2 ðsÞ

This equation can be obtained by taking the s-domain equivalent of the model in Figures 5.6.1(a) or
5.6.2(a) and setting up the mesh equation for it. That is why the models are called the ‘equivalents’ of
coupled circuits.
Figure 5.7.1(a) shows a circuit containing a pair of coupled coils with both dots marked on the same
side. The mesh equation for its s-domain equivalent can be set up with the coupled coils replaced by the
model in Figure 5.6.2(a) (with the upper one of the double signs before M), as depicted in Figure 5.7.1(b)
as
    
R1 þ sL1 þsM I1 ðsÞ Vi ðsÞ þ L1 i1 ð0Þ þ M i2 ð0Þ
¼ ð5:14Þ
þsM R2 þ sL2 þ 1=ðsC2 Þ I2 ðsÞ L2 i2 ð0Þ þ M i1 ð0Þ

where the signs of the mutual inductance terms having M are positive since both currents I1 ðsÞ and I2 ðsÞ
enter the dotted terminals of coil 1 and coil 2, respectively.
230 Chapter 5 Magnetically Coupled Circuits

Figure 5.7.1 A circuit containing a pair of coupled coils and its s-domain equivalent

Figure 5.7.2(a) shows a circuit containing a pair of magnetically coupled coils with the two dots
marked on opposite sides. The mesh equation for its s-domain equivalent can be set up with the coupled
coils replaced by the model in Figure 5.6.2(a) (with the lower one of the double signs before M) as
    
R1 þ sL1 sM I1 ðsÞ Vi ðsÞ þ L1 i1 ð0Þ  M i2 ð0Þ
¼ ð5:15Þ
sM R2 þ sL2 þ 1=ðsC2 Þ I2 ðsÞ L2 i2 ð0Þ  M i1 ð0Þ

where the signs of the mutual inductance terms having M are negative since one current I1 ðsÞ enters the
dotted terminal of coil 1 and the other current I2 ðsÞ enters the undotted terminal of coil 2. This is the same
as obtained by negating the terms in I2 ðsÞ and i2 ð0Þ in Equation (5.14) (see Figure 5.7.2(b)):

    
R1 þ sL1 þsM I1 ðsÞ Vi ðsÞ þ L1 i1 ð0Þ  M i2 ð0Þ
¼
þsM R2 þ sL2 þ 1=ðsC2 Þ I2 ðsÞ L2 i2 ð0Þ þ M i1 ð0Þ

It is implied that switching the winding direction of one coil has the same effect as switching the
reference direction of the current through the secondary coil (on the load side).

[Remark 5.2] Relative Winding and Current Reference Directions versus the Sign of Mutual Inductance
Consider the four quantities, i.e. the current reference directions and the winding directions (described
by the dot convention) of the two coils. If two or four of them are changed, it makes no difference in the
circuit equations for the coupled coils, but if one or three of them are changed, the sign of the terms
involving the mutual inductance will be reversed.

Figure 5.7.2 The same circuit as that of Figure 5.7.1(a), but with different winding direction or current reference
direction
5.4 Equivalent Models of Magnetically Coupled Coils 231

(Example 5.1) Mesh Analysis and Simulation of a Circuit Containing Coupled Coils

(a) Consider the circuit of Figure 5.8.1(a) in which the switch is closed at t ¼ 0 when the initial
conditions are iL1 ð0Þ ¼ 10=ð1 þ 1Þ ¼ 5 A and iL2 ð0Þ ¼ 0 A. Find iL1 ðtÞ and iL2 ðtÞ for t  0.
Like Equation (5.15), the mesh equation can be written and solved as
" #" # " # " #
1 þ 2s 4s I1 ðsÞ L1 i1 ð0Þ  M i2 ð0Þ 25
¼ ¼ ðE5:1:1Þ
4s 1 þ 8s þ 1=ð2sÞ I2 ðsÞ L2 i2 ð0Þ  M i1 ð0Þ 4  5
" # " #" # " #
I1 ðsÞ 1 1 þ 8s þ 1=ð2sÞ 4s 10 1 10s þ 5
¼ ¼
I2 ðsÞ 10s þ 2 þ 1=ð2sÞ 4s 1 þ 2s 20 20s2 þ 4s þ 1 20s
" #
1 ðs þ 1=10Þ þ 2ð1=5Þ
¼ 2 2
ðs þ 1=10Þ þ ð1=5Þ 2ðs þ 1=10Þ þ ð1=5Þ
" # " #
i1 ðtÞ et=10 ½cosðt=5Þ þ 2 sinðt=5Þ
¼ us ðtÞ ½A ðE5:1:2Þ
i2 ðtÞ et=10 ½2 cosðt=5Þ þ sinðt=5Þ

Figure 5.8.1 Circuits containing a pair of coupled coils for Example 5.1
232 Chapter 5 Magnetically Coupled Circuits

The following MATLAB program cir05e01a.m can be run to obtain the same results and plot
them for the time interval ½0; 100 s, as depicted in Figure 5.8.2(a):
>>cir05e01a
i1 ¼ exp(1/10*t)*cos(1/5*t) þ 2*exp(1/10*t)*sin(1/5*t)
i2 ¼ 2*exp(1/10*t)*cos(1/5*t) þ exp(1/10*t)*sin(1/5*t)

%cir05e01a.m
% To solve a circuit with magnetically coupled coils and SW (Ex 5.1a)
clear, clf
syms s
Rs¼1; R1¼1; R2¼1; L1¼2; L2¼8; M¼4; C2¼2;
i10¼5; i20¼0; % Initial conditions
Zs¼[R1þs*L1 s*M; s*M R2þ1/s/C2þs*L2]
Vs¼[L1*i10-M*i20; L2*i20-M*i10]
Is¼Zs\Vs % solution of Eq.(E5.1.1)
i1¼ilaplace(Is(1)), i2¼ilaplace(Is(2))
t0¼0; tf¼100; N¼500; tt¼t0þ [0:N]*(tf-t0)/N;
for n¼1:length(tt)
t¼tt(n); i1t(n)¼eval(i1); i2t(n)¼eval(i2);
end
subplot(221), plot(tt,i1t, tt,i2t), axis([0 100 -2 1.5])

(b) Consider the circuit of Figure 5.8.1(b) in which the switch is open at t ¼ 0 with zero initial
conditions iL1 ð0Þ ¼ 0 A and iL2 ð0Þ ¼ 0 A. Find iL1 ðtÞ and iL2 ðtÞ for t  0.
Like Equation (5.15), the mesh equation can be written and solved as

" #" # " # " #


1 þ 2s 4s I1 ðsÞ Vi ðsÞ 10=s
¼ ¼ ðE5:1:3Þ
4s 1 þ 8s þ 1=ð2sÞ I2 ðsÞ 0 0
" # " #" #
I1 ðsÞ 1 1 þ 8s þ 1=ð2sÞ 4s 10=s
¼
I2 ðsÞ 10s þ 2 þ 1=ð2sÞ 4s 1 þ 2s 0
2 3
10 2ðs þ 1=10Þ þ 4ð1=5Þ
" 6 s # 7
1 8s2 þ s þ 1=2 6 ðs þ 1=10Þ2 þ ð1=5Þ2 7
¼ 2 2
¼6 6
7
7
s½ðs þ 1=10Þ þ ð1=5Þ  4s2 4 4ðs þ 1=10Þ  2ð1=5Þ 5
2 2
ðs þ 1=10Þ þ ð1=5Þ
" # " t=10
#
i1 ðtÞ 10  e ½2 cosðt=5Þ þ 4 sinðt=5Þ ðE5:1:4Þ
¼ us ðtÞ ½A
i2 ðtÞ t=10
e ½4 cosðt=5Þ  2 sinðt=5Þ

MATLAB may be used to obtain the same results and plot them for the time interval ½0; 100 s, as
depicted in Figure 5.8.2(b):
>>cir05e01b
i1 ¼ 2*exp(1/10*t)*cos(1/5*t)4*exp(1/10*t)*sin(1/5*t)þ10
i2 ¼ 4*exp(1/10*t)*cos(1/5*t)2*exp(1/10*t)*sin(1/5*t)
5.4 Equivalent Models of Magnetically Coupled Coils 233

Figure 5.8.2 MATLAB analysis and PSpice simulation results for Example 5.1

(c) Consider the circuit of Figure 5.8.1(c), where a sinusoidal voltage source vi ðtÞ ¼ 10 sinð! tÞ [V]
with ! ¼ 200 ¼ 2f ½rad=s ðf ¼ 200=2 ’ 31:83 HzÞ is applied at t ¼ 0 when the initial condi-
tions are zero, i.e. iL1 ð0Þ ¼ 0 A and iL2 ð0Þ ¼ 0 A. Find iL1 ðtÞ and iL2 ðtÞ for t  0.
Like Equation (5.15), the mesh equation can be written and solved as
    
1 þ 0:002s 0:004s I1 ðsÞ Vi ðsÞ ¼ 2000=ðs2 þ 2002 Þ
¼ ðE5:1:5Þ
0:004s 1 þ 0:008s þ 103 =ð2sÞ I2 ðsÞ 0
Even with such unrealistically simple values of the parameters, the computation involved in
solving this equation and taking the inverse Laplace transform to obtain iL1 ðtÞ and iL2 ðtÞ seems
to be quite involved and there might be a need to resort to MATLAB. The MATLAB
program cir05e01c.m that follows is run to get the following results and plot them as depicted
in Figure 5.8.2(c):
234 Chapter 5 Magnetically Coupled Circuits

>>cir05e01c
52 98 52 64
i1¼ cos(200t) þ sin(200t) þ exp(-100t)cos(200t) þ exp(100t)sin(200t)
17 17 17 17
64 16 64 52
i2 ¼  cos(200t)   sin(200t)  exp(100t)cos(200t) þ exp(100t)sin(200t)
17 17 17 17

%cir05e01c.m
% To solve a circuit with coupled coils & a sinusoidal input (Ex 5.1c)
clear, clf
syms s
Rs¼0; R1¼1; R2¼1; L1¼0.002; L2¼0.008; M ¼0.004; C2¼0.002;
Zs¼ [RsþR1þs*L1 s*M; s*M R2þ1/s/C2þs*L2]
w¼200; Vs¼[10*w/(s^2þw^2); 0]
Is¼Zs\Vs % Eq.(E5.1.5)
i1¼ilaplace(Is(1)); i2¼ilaplace(Is(2)); pretty(i1), pretty(i2)
t0¼0; tf¼0.1; N¼500; tt¼t0 þ[0:N]*(tf-t0)/N;
for n¼1:length(tt)
t¼tt(n); i1t(n)¼eval(i1); i2t(n)¼eval(i2);
end
vit ¼ 10*sin(w*tt);
subplot(223), plot(tt,vit, tt,i1t, tt,i2t)

Note. The circuit diagrams of Figure 5.8.1 are PSpice schematics and the PSpice simulation results are
depicted side by side with the graphs obtained by using MATLAB in Figure 5.8.2.
Note. The coupled coils in Figure 5.8.1 are connected via a dummy resistor of extra-large resistance as required by the
PSpice rule that every node should have a DC path to the ground (see Remark H.2 in Appendix H).
Note. Figure 5.8.2 illustrates that the currents in the coupled coils may change instantaneously at t ¼ 0, which
is a surprising violation of the continuity rule of inductor currents. See Problem 5.12 for details.

5.4.2 -Equivalent Circuit


Equation (5.13) with zero initial conditions can be solved to give the expression of ½I1 ðsÞ I2 ðsÞ in terms
of ½V1 ðsÞ V2 ðsÞ as

    
I1 ðsÞ 1 sL2 sM V1 ðsÞ
¼
I2 ðsÞ s2 ðL1 L2  M 2 Þ sM sL1 V2 ðsÞ
  
1 L2  M  M M V1 ðsÞ
¼
sðL1 L2  M 2 Þ M L1  M  M V2 ðsÞ
    
I1 ðsÞ 1=ðsLa Þ þ 1=ðsLc Þ 1=ðsLc Þ V1 ðsÞ
¼ ð5:16Þ
I2 ðsÞ 1=ðsLc Þ 1=ðsLb Þ þ 1=ðsLc Þ V2 ðsÞ
L1 L2  M 2 L1 L2  M 2 L1 L2  M 2
with La ¼ ; Lb ¼ ; and Lc ¼ ð5:17Þ
L2  M L1  M M

This node equation directly corresponds to the -model for a pair of two magnetically coupled coils in
Figure 5.6.2(b), which can be used to set up the node equation for circuits containing coupled coils. Note
5.4 Equivalent Models of Magnetically Coupled Coils 235

Figure 5.9 A circuit containing a pair of coupled coils

that the sign of M should be negative or positive depending on whether the dots denoting the relative
winding directions of two coils are on the same or opposite sides.
For confirmation and practice, set up the node equation for the circuit in Figure 5.9(a). First, regarding
the two coil currents i1 and i2 as given, KCL is applied to nodes 1 and 2 to write

I1 ðsÞ þ ðG1 þ G12 ÞV1 ðsÞ  G12 V2 ðsÞ ¼ Ii ðsÞ


I2 ðsÞ  G12 V1 ðsÞ þ ðG2 þ G12 ÞV2 ðsÞ ¼ 0

Substituting Equation (5.16) for I1 ðsÞ and I2 ðsÞ into this equation yields

2 3
L2 M
G1 þ G12 þ G 12     
6 sðL1 L2  M 2 Þ 7
6 sðL1 L2  M 2 Þ 7 V1 ðsÞ ¼ Ii ðsÞ ð5:18Þ
4 M L1 5 V2 ðsÞ 0
G12  G2 þ G12 þ
sðL1 L2  M 2 Þ sðL1 L2  M 2 Þ

This is identical to the node equation for the circuit of Figure 5.9(b), in which the pair of two coupled
circuits is replaced by the -model in Figure 5.6.2(b) consisting of three uncoupled coils.

(Example 5.2) Node Analysis and Simulation of a Circuit Containing Coupled Circuits
Consider the circuit in Figure 5.10.1(a1) where R1 ¼ 10 , R12 ¼ 10 , R2 ¼ 5 , L1 ¼ 1 H,
L2 ¼ 2 H, M ¼ 1 H, and the current source of 10 A is applied at t ¼ 0 when the initial conditions
are zero; i.e. iL1 ð0Þ ¼ 0 A and iL2 ð0Þ ¼ 0 A. Find v1 ðtÞ and v2 ðtÞ for t  0.
From Equation (5.18) with G1 ¼ 1=10 S, G12 ¼ 1=10 S, G2 ¼ 1=5 S, L1 ¼ 1 H, L2 ¼ 2 H,
M ¼ 1 H, and Ii ðsÞ ¼ 10=s ½A, the node equation can be written and solved as

" #" # " #


0:1 þ 0:1 þ 2=s 0:1  1=s V1 ðsÞ 10=s
¼ ðE5:2:1Þ
0:1  1=s 0:1 þ 0:2 þ 1=s V2 ðsÞ 0
" # " #" #
V1 ðsÞ 1 0:3 þ 1=s 0:1 þ 1=s 10=s
¼
V2 ðsÞ 0:05 þ 0:06=s þ 1=s2 0:1 þ 1=s 0:2 þ 2=s 0
" # " #
20 3s þ 10 10=ðs þ 2Þ þ 50=ðs þ 10Þ
¼ ¼
ðs þ 2Þðs þ 10Þ s þ 10 20=ðs þ 2Þ
" # " #
v1 ðtÞ 10 e2t þ 50 e10t
¼ us ðtÞ ½V ðE5:2:2Þ
v2 ðtÞ 20 e2t
236 Chapter 5 Magnetically Coupled Circuits

Figure 5.10.1 Simulation of a circuit containing a pair of coupled coils

Readers are invited to compose a MATLAB program named, say, cir05e02.m to get the same
results and plot them for the time interval ½0; 1s, as depicted in Figure 5.10.1(a2):

>>cir05e02
v1 ¼ 20*exp(6*t)*(3*cosh(4*t)  2*sinh(4*t))
v2 ¼ 20*exp(2*t)
Note. Note that cosh 4t ¼ ðe4t þ e4t Þ=2 and sinh 4t ¼ ðe4t  e4t Þ=2.)

Figures 5.10.1(a1) and (b1) are PSpice schematics themselves, where the latter one has the -model
for the pair of coupled coils. Since the two coupled coils are interconnected via R12 , unlike those in
Figure 5.8.1, they do not have to be connected via a dummy resistor of extra-large resistance, but
have to be connected via a dummy resistor of extra-small resistance, or each of the two coils can
be grounded separately. Especially when the two coils L10 and L20 are shorted between their lower
parts in the PSpice schematic in Figure 5.10.1(b1), it will cause a run-time error since any
loop consisting of inductors only ðL10 –L12 –L20 Þ is rejected by PSpice (Remark H.2(3)). Figure
5.10.2(a) shows the Property Editor spreadsheet for the pair of coupled coils in Figure 5.10.1(a1),
where the two inductances are set to L1 ¼ 1 H and L2 ¼ 2 H, respectively and the coefficient of
coupling is set to

M 1
k ¼ pffiffiffiffiffiffiffiffiffiffi ¼ pffiffiffi ’ 0:7071 ðE5:2:3Þ
L1 L2 2

Figure 5.10.2(b) shows the Simulation Settings dialog box, where the square box before SKIPBP
(Skip the initial transient bias point calculation) is checked.
5.5 Ideal Transformer 237

Figure 5.10.2 Property Editor spreadsheet and simulation setting dialog box for PSpice simulation of
Figure 5.10.1(a1)

5.5 Ideal Transformer


The conditions for a pair of coupled coils to be an ideal transformer are as follows:

1. The two coils are perfectly coupled with the unity coefficient of coupling k ¼ 1:

ð5:7Þ pffiffiffiffiffiffiffiffiffiffi
M ¼ L1 L2 ð5:19aÞ
with k¼1

2. The permeance P of the magnetic core around which the two coils are wound is 1 or, equivalently,
the reluctance is R ¼ 1=P ¼ 0 so that the magnetomotive force (mmf) around the magnetic circuit is
zero however large the flux  may be:
ð5:1Þ i2 N1
N1 i1  N2 i2 ¼ R ¼ 0; ¼ ð5:19bÞ
i1 N2
238 Chapter 5 Magnetically Coupled Circuits

3. No energy is stored or dissipated in the two coils, which means that all the power received at one
(primary or source) side is instantly transferred to the other (secondary or load) side:

v2 i1
pðtÞ ¼ v1 i1 þ v2 i2 ¼ 0; v1 i1 ¼ v2 i2 ; ¼ ð5:19cÞ
v1 i2

Noting from Equation (5.3) that the self-inductance of each coil wound around a common core (with
permeance P) is proportional to the square of the number of turns, we have

N22 ð5:19aÞ pffiffiffiffiffiffiffiffiffiffi N2 N2


L2 ¼ L1 ; M ¼ L1 L2 ¼ L1 ; and L2 ¼ M ð5:20Þ
N12 N1 N1

so that the voltage–current relationship of an ideal transformer can be written as

ð5:9aÞ di1 di2


v1 ðtÞ ¼ L1 M
dt dt
ð5:9bÞ N2 di1 N2 di2 N2
v2 ðtÞ ¼  L1 þ M ¼  v1 ðtÞ
ð5:20Þ N1 dt N1 dt N1

This relationship between the primary–secondary voltages can be obtained by substituting Equation
(5.19b) into Equation (5.19c). It is implied by this result and Equation (5.19c) that the primary–
secondary voltages of an ideal transformer are proportional to the number of turns of coil winding,
while the primary–secondary currents are inversely proportional to the number of turns of coil winding,
which can be summarized as below, where the turns ratio is defined as a ¼ N1 =N2 .

Relationships between the primary–secondary voltages and currents of an ideal transformer:


v2 N2 1
¼ ¼ ð5:21aÞ
v1 N1 a
i2 N1 N1
¼ ¼ a with the turns ratio a ¼ ð5:21bÞ
i1 N2 N2

Note. The upper/lower signs apply for the case of positive/negative mutual inductance, respectively.

These relationships between the primary–secondary voltages and currents can be modeled by the circuits
containing dependent sources, as shown in Figures 5.11(a) and (b).
Based on the voltage–current transformation properties (5.21a) and (5.21b), another important
property of impedance transformation (or multiplication or scaling) possessed by an ideal transformer

Figure 5.11 Dependent source models for an ideal transformer


5.5 Ideal Transformer 239

will be derived. For this purpose, consider the ideal transformer circuit of Figure 5.12(a) in which the
voltage–current relationship of the load is written as

V2 ðsÞ ¼ ZL ðsÞIL ðsÞ ¼ ZL ðsÞI2 ðsÞ

With Equation (5.21b) this can be substituted into Equation (5.21a) to get

 
ð5:21aÞ N1 N1 ð5:21bÞ N1 N1
V1 ðsÞ ¼ V2 ðsÞ ¼ ½ZL ðsÞI2 ðsÞ ¼  ZL ðsÞ  I1 ðsÞ
N2 N2 N2 N2
 2
N1
V1 ðsÞ ¼ ZL ðsÞI1 ðsÞ ð5:22Þ
N2

Thus the secondary (load) impedance reflected to the primary (source) side is

 2
V1 ðsÞ N1 N1
Z12 ðsÞ ¼ ¼ ZL ðsÞ ¼ a2 ZL ðsÞ with a ¼ ð5:23Þ
I1 ðsÞ N2 N2

This reflected impedance implies that, from the primary (source) side, the secondary (load) impedance is
seen to be multiplied by the squared turns ratio ða2 Þ. This property is not only helpful in understanding
the basic function of a transformer but is also useful in the realization of maximum power transfer or
impedance matching, which will be discussed in Section 6.7.
Similarly, referring to Figure 5.12(b), use can be made of Equations (5.21a) and (5.21b) to find the
Thevenin equivalent of the transformer circuit seen from the secondary (load) side as

 
ð5:21aÞ N2 N2 ð5:21bÞ N2 N2
V2 ðsÞ ¼  V1 ðsÞ ¼  ½Vs ðsÞ  Zs ðsÞI1 ðsÞ ¼  Vs ðsÞ  Zs ðsÞ I2 ðsÞ
N1 N1 N1 N1
 2
N2 N2 1 1
V2 ðsÞ ¼  Vs ðsÞ þ Zs ðsÞI2 ðsÞ ¼  Vs ðsÞ þ 2 Zs ðsÞI2 ðsÞ ð5:24Þ
N1 N1 a a

This implies that from the secondary (load) side, the source voltage and the primary (source) impedance
are seen to be multiplied by the reverse turns ratio 1=a and the reverse squared turns ratio 1=a2 ,
respectively. This property can be used to eliminate the transformer in order to simplify the analysis
of a transformer circuit. However, it does not apply in the case where there is some external connection
between the two coils.

Figure 5.12 Impedance transformation (multiplication) by an ideal transformer


240 Chapter 5 Magnetically Coupled Circuits

(Example 5.3) Electric Power Transmission with High Voltage Using Transformers
Consider the circuit of Figure 5.12(a) in which the turns ratio of the ideal transformer is
a ¼ N1 =N2 ¼ 30=1, the primary voltage is V1 ¼ 6600 V, and the load impedance is a resistance of
ZL ¼ 22 . The secondary voltage/current and the primary current are

ð5:21aÞ N2 1 V2 220
V2 ¼ V1 ¼ 6600 ¼ 220 V; I2 ¼ IL ¼  ¼ ¼ 10 A
N1 30 ZL 22

and
ð5:21bÞ N2 1
I1 ¼  I2 ¼ A ðE5:3:1Þ
N1 3

respectively. The load impedance reflected to the primary (source) side is

V1 ð5:23Þ 2
Z12 ¼ ¼ a ZL ¼ 302  22 ¼ 19 800  ðE5:3:2Þ
I1

This implies that despite the high primary voltage, the load impedance seen from the primary side is
magnified a2 ¼ 900 times the original value, so that the primary current is merely 1/3 A, much less
than the secondary (load) current IL ¼ V2 =ZL ¼ 10 A.
Note. If the primary current flows through a long transmission line from the generator, this small current I1 will
be good for decreasing the transmission loss as well as the voltage drop. This is why the high voltage
transmission is adopted for large power systems.

5.6 Linear Transformer


Figure 5.13 shows a more realistic model for a transformer, which contains the internal coil resistances.
To find the reflected impedance, the voltage gain, and the current gain, all the initial conditions
are neglected, the mesh equation is set up in the primary/secondary currents I1 ðsÞ and I2 ðsÞ, and it is
solved as
½Zs ðsÞ þ R1 þ sL1 I1 ðsÞ þ s M I2 ðsÞ ¼ Vs ðsÞ
s M I1 ðsÞ þ ½R2 þ sL2 þ ZL ðsÞI2 ðsÞ ¼ 0
    
Z11 ðsÞ sM I1 ðsÞ Vs ðsÞ
¼
sM Z22 ðsÞ I2 ðsÞ 0
with Z11 ðsÞ ¼ Zs ðsÞ þ R1 þ sL1 and Z22 ðsÞ ¼ R2 þ sL2 þ ZL ðsÞ
      
I1 ðsÞ 1 Z22 ðsÞ sM Vs ðsÞ Vs ðsÞ Z22 ðsÞ
¼ ¼ ð5:25Þ
I2 ðsÞ  sM Z11 ðsÞ 0 Z11 ðsÞZ22 ðsÞ  s2 M 2 sM

Figure 5.13 The s-domain linear transformer model


5.7 Autotransformers 241

Thus the input impedance of the overall circuit seen from the source is

Vs ðsÞ Z11 ðsÞZ22 ðsÞ  s2 M 2 s2 M 2


Zin ðsÞ ¼ ¼ ¼ Z11 ðsÞ 
I1 ðsÞ Z22 ðsÞ Z22 ðsÞ
s2 M 2
¼ Zs ðsÞ þ R1 þ sL1  ð5:26Þ
sL2 þ R2 þ ZL ðsÞ

Neglecting the source impedance Zs ðsÞ and the internal resistances R1 and R2 and substituting the
conditions of an ideal transformer
ð5:19aÞ pffiffiffiffiffiffiffiffiffiffi
L1 ¼ N12 P; L2 ¼ N22 PðP ¼ 1Þ; and M ¼ L1 L2 ð5:27Þ
with k¼1

into Equation (5.26) yields the load impedance reflected to the primary (source) side as
 2
s2 M 2 L1 L2 ¼M 2 sL1 ZL ðsÞ L2 ZL L1 ZL ðsÞ N1
Z12 ðsÞ ¼ sL1  ¼ ’ ¼ ZL ðsÞ ð5:28Þ
sL2 þ ZL ðsÞ sL2 þ ZL ðsÞ L2 N2

which agrees with Equation (5.23) for an ideal transformer.


The voltage gain, i.e. the ratio of the secondary (load) voltage to the source voltage, is

V2 ðsÞ I2 ZL ðsÞ ð5:25Þ sM ZL ðsÞ


Av ¼ ¼ ¼
Vs ðsÞ Vs ðsÞ Z11 ðsÞZ22 ðsÞ  s2 M 2
sM ZL ðsÞ
¼ ð5:29Þ
½Zs ðsÞ þ R1 þ sL1 ½R2 þ sL2 þ ZL ðsÞ  s2 M 2

and the current gain, i.e. the ratio of the secondary (load) current to the primary (source) current is

I2 ðsÞ ð5:25Þ sM sM


Ai ¼ ¼ ¼ ð5:30Þ
I1 ðsÞ Z22 ðsÞ R2 þ sL2 þ ZL ðsÞ

Neglecting the source impedance Zs ðsÞ and the internal resistances R1 and R2 and substituting the ideal
transformer conditions (5.27) into these two equations, (5.29) and (5.30), yields

V2 ðsÞ sM ZL ðsÞ N2
Av ¼ ’ ¼ ð5:31aÞ
Vs ðsÞ sL1 ZL ðsÞ N1
I2 ðsÞ sM N1
Ai ¼ ’ ¼ ð5:31bÞ
I1 ðsÞ sL2 N2

which agree with the primary–secondary voltage and current relationship, (5.21a) and (5.21b), for an
ideal transformer.

5.7 Autotransformers
Figures 5.14(a) and 5.15(a) show a step-up autotransformer and a step-down one, respectively, in which
the primary and secondary coils have some or all windings in common and the turns ratio depends on the
position of the connection point called a tap. Compared with a conventional two-winding transformer
with the same turns ratio, an autotransformer is lighter, smaller, and less costly because it requires both
242 Chapter 5 Magnetically Coupled Circuits

Figure 5.14 A step-up autotransformer and its s-domain equivalent

fewer windings and a smaller core. On the other hand, it does not have the function of electrical isolation
to reduce the risk of shock hazard or to remove the DC influence of one side on the other.
The coupled coils in Figure 5.14(a) can be replaced by the T-equivalent (Figure 5.6.2(a)), as depicted
in Figure 5.14(b), and the mesh equation set up as

    
sL1 sðL1 þ MÞ I1 ðsÞ V1 ðsÞ
¼ ð5:32Þ
sðL1 þ MÞ sðL1 þ 2M þ L2 Þ þ ZL ðsÞ I2 ðsÞ 0

which is solved to find the primary/secondary currents, the voltage gain, and the current gain as

    
I1 ðsÞ 1 sðL1 þ 2M þ L2 Þ þ ZL ðsÞ sðL1 þ MÞ V1 ðsÞ
¼
I2 ðsÞ sðL1 L2  M 2 Þ þ sL1 ZL ðsÞ sðL1 þ MÞ sL1 0
 
V1 ðsÞ sðL1 þ 2M þ L2 Þ þ ZL ðsÞ
¼ ð5:33Þ
sðL1 L2  M 2 Þ þ sL1 ZL ðsÞ sðL1 þ MÞ
I2 ðsÞ ð5:33Þ sðL1 þ MÞ
Ai ¼ ¼ ð5:34Þ
I1 ðsÞ sðL1 þ 2M þ L2 Þ þ ZL ðsÞ
V2 ðsÞ I2 ðsÞZL ðsÞ ð5:33Þ sðL1 þ MÞZL ðsÞ
Av ¼ ¼ ¼ ð5:35Þ
V1 ðsÞ V1 ðsÞ sðL1 L2  MÞ þ sL1 ZL ðsÞ

Figure 5.15 A step-down autotransformer and its s-domain equivalent


Problems 243

Substituting the ideal transformer conditions (5.27) into these two equations yields
pffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffi
I2 ðsÞ ð5:34Þ with ð5:27Þ L1 þ L1 L2 L1 N1
Ai ¼ ¼ pffiffiffiffiffiffiffiffiffiffi pffiffiffi ¼ pffiffiffiffiffi pffiffiffi ¼ ð5:36Þ
I1 ðsÞ L1 þ 2 L1 L2 þ L2 L1 þ L2 N1 þ N2
V2 ðsÞ ð5:35Þ with ð5:27Þ L1 þ M N12 þ N1 N2 N1 þ N2
Av ¼ ¼ ¼ ¼ ð5:37Þ
V1 ðsÞ L1 N12 N1
Likewise, the coupled coils in Figure 5.15(a) can be replaced by the T-equivalent (Figure 5.6.2(a)), as
depicted in Figure 5.15(b), and the mesh equation is set up as
    
sðL1 þ 2M þ L2 Þ sðL1 þ MÞ I1 ðsÞ V1 ðsÞ
¼ ð5:38Þ
sðL1 þ MÞ sL1 þ ZL ðsÞ I2 ðsÞ 0
which is solved to find the primary/secondary currents, the voltage gain, and the current gain as
" # " #" #
I1 ðsÞ 1 sL1 þ ZL ðsÞ sðL1 þ MÞ V1 ðsÞ
¼
I2 ðsÞ  sðL1 þ MÞ sðL1 þ 2M þ L2 Þ 0
" #
V1 ðsÞ sL1 þ ZL ðsÞ
¼ 2 ð5:39Þ
s ðL1 L2  M 2 Þ þ sðL1 þ 2M þ L2 ÞZL ðsÞ sðL1 þ MÞ

I2 ðsÞ ð5:39Þ sðL1 þ MÞ


Ai ¼ ¼ ð5:40Þ
I1 ðsÞ sL1 þ ZL ðsÞ
V2 ðsÞ I2 ðsÞZL ðsÞ ð5:39Þ sðL1 þ MÞZL ðsÞ
Av ¼ ¼ ¼ 2 ð5:41Þ
V1 ðsÞ V1 ðsÞ s ðL1 L2  M 2 Þ þ sðL1 þ 2M þ L2 ÞZL ðsÞ
Substituting the ideal transformer conditions (5.27) into these two equations yields
pffiffiffiffiffiffiffiffiffiffi
I2 ðsÞ ð5:40Þ with ð5:27Þ L1 þ L1 L2 N12 þ N1 N2 N1 þ N2
Ai ¼ ¼ ¼ ¼ ð5:42Þ
I1 ðsÞ L1 N12 N1
V2 ðsÞ ð5:41Þ with ð5:27Þ L1 þ M N 2 þ N1 N2 N1
Av ¼ ¼ ¼ 2 1 2
¼ ð5:43Þ
V1 ðsÞ L1 þ 2M þ L2 N1 þ 2N1 N2 þ N2 N1 þ N2

Note. A failure of insulation for windings of an autotransformer may cause the full source voltage/current to be applied
to the load side.

Problems
5.1 Series Connections of Coupled Coils

Figure P5.1 Series connections of two coupled coils

Вам также может понравиться