Sei sulla pagina 1di 50

ENGN4625/6625 2014 eBrick 1 - 1

Mathematics for Power Circuit


Analysis
Contents
Mathematics for Power Circuit Analysis ..........................................................................................1
1 Complex notation & phasors .......................................................................................................3
1.1 Phasors ..................................................................................................................................3
1.1.1 Phase, “Lagging” and “Leading” ...................................................................................3
1.1.2 Phasors are RMS ............................................................................................................4
2 Power Calculations and RMS ......................................................................................................4
2.1 Power and Average Power ....................................................................................................4
2.1.1 RMS and Crest Factor ....................................................................................................4
2.1.2 RSS ................................................................................................................................5
2.1.3 Non-Linear Power Calculations .....................................................................................6
2.2 Power Factor, Complex Power .............................................................................................6
2.2.1 Power in an AC circuit ...................................................................................................6
2.2.2 Complex Power, Reactive Power (VAr), Apparent Power (VA) ..................................7
2.2.3 Example calculation of S, P, Q: .....................................................................................7
2.2.4 Why is complex power S = V × I* ? .............................................................................8
2.2.5 Example: Obtaining Z, given power and power factor .................................................8
2.2.6 Extension of Power Factor to non-Sinusoidal Waveforms: ...........................................9
2.2.7 Power Factor Correction ................................................................................................9
3 Three Phase Power .....................................................................................................................11
3.1.1 First - Why do we use AC? ..........................................................................................11
3.1.2 Why three phase? .........................................................................................................11
3.1.3 Three phase Rotation Conventions ..............................................................................11
3.1.4 Currents and voltages in balanced three phase networks.............................................12
3.1.5 Three phase circuit connections – Wye and Delta .....................................................13
3.2 Balanced and Unbalanced loads ..........................................................................................13
3.2.1 The h operator and the H matrix ..................................................................................13
3.2.2 Example Calculation with neutral intact – by inspection. ...........................................14
3.2.3 Three wire connection (no neutral) ..............................................................................14
3.2.4 Power in polyphase (eg. three phase) circuits .............................................................14

E. Franklin, B. Blackwell p1/50


ENGN4625/6625 2014 eBrick 1 - 2

4 The Nodal Admittance Matrix method ......................................................................................15


4.1 Determining the NAM by inspection .................................................................................15
4.1.1 Example Calculation ....................................................................................................17
4.2 Star-Delta transformation via the NAM ..............................................................................18
5 Power Transmission ...................................................................................................................19
5.1 Impedances of transmission lines ........................................................................................20
5.2 Transmission line models ....................................................................................................21
5.3 Propagation on transmission lines .......................................................................................22
6 Coordinate Transformations ......................................................................................................24
6.1 Introduction .........................................................................................................................24
6.2 Power Invariance .................................................................................................................24
6.3 The Symmetric Component Transform...............................................................................25
6.3.1 Example calculation of symmetric components ..........................................................27
6.3.2 Simplification of a general, balanced static impedance ...............................................27
6.3.3 Simplification of a general, balanced AC machine impedance ...................................28
6.3.4 Important Properties of the Symmetric Component Transform ..................................28
6.3.5 Applications of the Symmetric Component Transform ...............................................28
6.3.6 Negative Sequence Detection in Motors......................................................................29
6.3.7 Line Fault Analysis ......................................................................................................29
6.4 Time domain analysis..........................................................................................................33
6.4.1 The Clarke or α, β, 0 Transform for time domain analysis ..........................................33
6.5 The Park’s or D,Q,0 Transform for rotating machines .......................................................35
7 Per Unit Notation .......................................................................................................................40
7.1.1 Example using per-unit notation: .................................................................................40
8 Fourier Series .............................................................................................................................42
8.1 Fourier series definition ......................................................................................................42
8.1.1 Symmetry .....................................................................................................................42
8.1.2 Example - fourier series coefficients for a square wave ..............................................43
8.2 Consideration of fourier analysis for power conversion circuits ........................................44
9 Computer Simulation – LTC Spice............................................................................................46
9.1.1 Hints for running LTC Spice .......................................................................................48
9.2 pSpice (pSp) and LTSpice (LT) hints and comparisons. ....................................................48
9.2.1 Accuracy: .....................................................................................................................49
9.2.2 Saving setup time in “Probe”: ......................................................................................49
9.2.3 Libraries: ......................................................................................................................49

E. Franklin, B. Blackwell p2/50


ENGN4625/6625 2014 eBrick 1 - 3

1 Complex notation & phasors

sinusoidal waveform. A sinusoid of arbitrary phase , = cos( + ) can be represented as


Complex numbers provide a means to conveniently encode both the phase and magnitude of a

the sum of two components, the component in phase with a reference signal (the real part), and the

Mathematically, this is described via Euler’s formula ( = cos + sin ) by,


component lagging by 90°, the quadrature component (which will be the imaginary part).

cos( + ) = [ ( )
+ ( )
] ( 1)

the first term ( )


alone, the second term of the result will just be the conjugate. Then just take
Noting that the second exponential term is the conjugate of the first, we can do all calculations with

the real part of the result to get the answer because:

cos( + ) = Re" ( )
# = Re$cos( + ) + sin( + )%

Consider for example the sum of two sinusoidal voltages & cos( + &) + ' cos( + ')

The sum of these voltages (or indeed any other combination of operations on sinusoidal signals)
could be done using trigonometric identities. However it is far simpler to use complex number
notation, so that the problem then reduces to the sum of complex numbers, ie:

& cos( + &) + ' cos( + ') =( " + #


( )) ( *)
& '

= ( "( &
) + '
* ) #

1.1 Phasors
Above, the variables were written in lower case as they represented explicit functions of time. In
phasor notation we assume an exponential time variation at the given frequency, and simply write
the complex number in capitals corresponding to the RMS value and the phase.

Phasors can be written using in rectangular co-ordinates or one of two notations in polar co-
ordinates, for example:

+& = |+& | cos & + |+& |sin & -. +& = |+& | ) /0 |+& |∠ &

1.1.1 Phase, “Lagging” and “Leading”


Phasors (3-2-3 Mohan) can bet viewed as rotating
vectors, and are assumed to rotate anti-clockwise in
time. We normally draw phasors at t=0, (a snapshot
in time) but if we draw a vector representation of Va Vb
Va
at three consecutive time intervals, it looks like Fig Va
ωt
1. A waveform which is delayed in time (see Vb,
dashed, red) is said to have a phase lag. Its complex Vb
t=0 t =T/3 t =2T/3
representation has a negative imaginary part, and the Vb Va
magnitude phase representation has a negative angle
(240∠-30 in the figure). Consider the current in an
Figure 1: VA Real and Imag parts at 3 consecutive
ideal inductor, which lags the voltage by 90°, so the times, VB lags VA
current phasor has a negative imaginary part, and a

E. Franklin, B. Blackwell p3/50


ENGN4625/6625 2014 eBrick 1 - 4

negative angle (∠-90). A consequence of this is that an inductive impedance (V/I) has a positive
imaginary part, as the imaginary part changes sign during division. You will learn that an inductor
has a lagging power factor (see inset box). We can re-use the same diagram if we re-use the phasor
Vb as the current in a lossy inductor (Z=R+jωL) with a voltage Va applied. The phasor orientation
shows that the current lags by about 30° so this is quite a lossy inductor.

For later: A lagging power factor (e.g. inductor) is deemed to be positive, just like the sign of the imaginary part
of its impedance, even though the current has a negative phase angle.

1.1.2 Phasors are RMS


In power engineering we usually use RMS (root mean square) phasors by default, so 240∠120°
means a cosine advanced by 120° (i.e. same as delayed by 240°) with amplitude 339.4V (240*√2).
Note that Spice (LTspice and pSpice) uses amplitudes for AC, because we are usually following
voltage with time, so you would need to enter 339.4 there. (It only matters if you plot
voltages/currents versus time.)

2 Power Calculations and RMS

2.1 Power and Average Power


Power is the product of VI (strictly speaking it is vi instantaneously). For 50Hz supply, the power

cos( ) ∗ cos( ) = + cos(2 ) ), but we are usually more concerned with the practical
in a purely resistive load therefore oscillates at 100Hz with amplitude between 0 and Pmax (since

terms like average power.

1 ;
< 4 > = 8 VI dt
T <
The average power <P> over interval T is ( 2)

We normally average over one cycle, so T=1/f.

First consider linear networks. If the load is linear (V=IR) and we know R, we can use V∝ I to
allow calculation from just one measurement (V or I), which is where (and only where) the concept
of RMS (Root Mean Square) is strictly applicable. Substituting V=IR,

1 ; R ;
< 4 > = 8 VI dt = 8 I dt
T < T <

2.1.1 RMS and Crest Factor


If we define the RMS current IRMS as the equivalent current which would cause the same heating
effect as the actual, time-varying current in a linear circuit, then

R ;
< 4 > = R. I@AB = 8 I dt
T <

E. Franklin, B. Blackwell p4/50


ENGN4625/6625 2014 eBrick 1 - 5

I@AB = CD E< I dtF


;
;
Hence, we obtain ( 3)

Crest factor for sine waves is of course the familiar √2. Crest factor for other common waveforms are
Furthermore, we define Crest Factor as Crest Factor = Ipeak/IRMS ( 4)

given in the table.

IAverage/IPeak Crest factor (Ipeak/IRMS)

√2
DC constant signal 1 1
Sine wave 0

√3
Square wave 0 1

√3
Triangle wave 0

1/√D
Sawtooth wave 0
Pulse train wave with duty cycle D D

We must remember that we assumed a linear circuit, so that V∝ I. Power can only be calculated
from RMS if this is true. i.e. RMS only makes sense in a linear circuit (eg resistive, inductive, &
capacitive only). If you make this mistake in a lab report, you lose 2 marks! See Non-Linear
Power below.

Please also note: despite what audio retailers say, “RMS power” doesn’t make sense either. This
will also cost you marks in a lab report. What they mean is the average power, evaluated for a long
time, calculated from (VRMS)2 / RLoad If you actually evaluated the RMS power, it wouldn’t even
have the units of power. [not Joules/sec but Joules/(second)3/2]

2.1.2 RSS

if they are independent (orthogonal). For example, if x(t) and y(t) are orthogonal (i.e. L E MN . =
A related concept is Root Sum Squares (RSS) which is a rule for adding RMS quantities if and only

0 ) then the RMS of M( ) + N( ) is sqrt(R + S ) , where X,Y are the RMS values of M( ), N( ).
This applies to harmonics, with T = UT< for U integral. You can simply add the squares and take
the square root. This is also true of sin and cosine. A “proof” is loosely (consider RMS squared)

RMS2 (x+y) = L
E(M + N) . = L E(M + 2MN + N^2) . = L E M . + L E N . + L E MN .

= RMS2(x) + RMS2(y) + 0

Another viewpoint is that RSS can be used to combine powers of different frequency components
because products of I and V have zero average unless frequencies of V and I match, causing real
power dissipation.

equivalent series of harmonics1 is given simply using RSS: = sqrt(1 + XY + ZY … ) = 1.1107.


Example of RSS: If the fundamental of a square wave has RMS value 1, then the RMS of the

For the case of a square wave1 whose amplitude is 1, the RMS is 4/π √(1 + XY + ZY …) × 1/√2 = 1.

wave of amplitude 1, sqrt( L E . ) = 1.


In this case it is actually much easier to calculate the integral of for a square wave. For square

1
note that the Fourier series for a square wave is 4/][sin( ) + 1/3 sin(3 ) + 1/5 sin(5 ) + … ].
E. Franklin, B. Blackwell p5/50
ENGN4625/6625 2014 eBrick 1 - 6

Ultimately, we must use ( 2) < 4 > = E< VI dt, to calculate average power for non-linear
;
2.1.3 Non-Linear Power Calculations

;
circuits (since the RMS shortcut cannot be used). This can be done on a digital oscilloscope with
taking care that the zeros are correct (especially important for a Hall current probe). However there
are two special cases which occur in the lab. If one of V or I is constant, it can be factored out of
the average:

– say V is constant (the input to a buck convertor).

then <P> = <V.I> = V<I>, ( 5)

which is just the DC meter reading of V times the DC meter reading of I. (the DC or direct current
value of a waveform is its average). Thus if you simply measure with a DC (NOT AC or RMS)
meter, AND you check with your oscilloscope that at least one of V or I is constant, then you have
the power!

2.2 Power Factor, Complex Power


A very important measure to characterise the efficiency of an electrical device (load) is the power
factor. In simple terms, it is the power consumed by a load, as a fraction of the power that would
be consumed by an ideal resistive load connected to a sinusoidal source, which draws the same
RMS current and voltage.

First we consider the power in a linear load of complex impedance Z, if I leads V by .

2.2.1 Power in an AC circuit


We go back to cos(ωt) notation and Equation ( 3) to see how to calculate AC power.

If = cos(_ ) -. ` = cos(_ + ) , then

a = ` = cos(_ ) cos(_ + ) = cos(_ )[cos(_ ) cos( ) − sin(_ ) sin( )]

We obtain

1 c 1 c1 1 1
< 4 > = 8 ` de = 8 f1 + cos(2 )gcos + sin(2 )sin . = cos
c 0 c 0 2 2 2

Which, if we use V and I (RMS values factor of √2 twice)

< 4 > = +h cos


Here, cos is called the Power Factor, with a value 1 for a perfect or resistive load, and
yields ( 6)

less for a reactive load. Because cos is an even function (same value regardless of the sign of
the angle), the fraction of power delivered is independent whether the phase is leading or lagging,
we need to add one of those two words to make it clear – e.g.

“power factor is 0.5 lagging”

describes a significantly inductive load with a phase lag of 60° - I lags V by 60°, so = -π/3.

Such a load heats the wiring and the transformers (and ultimately the generator) supplying it twice
as much as an ideal load, so it must be compensated for by “power factor correction”.

E. Franklin, B. Blackwell p6/50


ENGN4625/6625 2014 eBrick 1 - 7

Careful with signs! A lagging power factor (e.g inductor) is deemed to be positive, just like the sign
of the imaginary part of its impedance, even though the current has a negative phase angle. The
maths is consistent, but the wording is tricky. Don’t assume lagging means negative or positive
without considering what quantity you are describing: a lagging power factor is given a positive
angle (corresponds to the complex power and impedance), but the current in that circuit (which
lags the voltage) has a negative angle and imaginary part.

2.2.2 Complex Power, Reactive Power (VAr), Apparent Power (VA)

A useful concept is complex power. i = +h∗, where h∗ is the complex conjugate of h ( 7)

The complex conjugate produces a real result for V and I in phase, for any complex V and I. The
real part of the complex power is the Real Power or average power as we have just calculated in (
6). The imaginary part is the Reactive Power, which corresponds to power oscillating in and out
of a circuit, and transferring no energy on the average. In an LC circuit the energy sloshes between
the inductor and capacitor, backwards and forwards. The only real power is a small amount which
is absorbed due to losses.

i = +h ∗ = |+||h| , where =∠V-∠I

= |+||h| cos + |+||h| sin ( 8)

Hence we write S = P + jQ where P is Real Power, Q is Reactive Power.

P = |+||h| cos and Q = |+||h| sin

The power factor is then the Real Power/ Apparent Power

4j = 4/|i| = cos ( 9)

where Apparent Power is the RMS voltage times the RMS current = |V| × |I| = |S|, the power you
would assume if you didn’t know the phase (quite a common situation).

Reactive power is measured in units of VAr (Volt-Amps-reactive), pronounced Vars, whereas the
units of apparent power are Volt-Amps or VA, pronounced Vee-Ay, not var.

2.2.3 Example calculation of S, P, Q:


An induction motor running on 240V draws a current of 5A at lagging by 60° (bad!). As usual in
power electronics, these are measured in RMS.

The apparent power |S| = |V||I| = 1200VA, the power (real power P) is 1200cos(60) = 600W, and
the reactive power is √3/2 × 1200 ~ 1039VAr.

Or S = P + jQ = VI* = 240∠0 × (5∠-60)* = 1200 [cos(]/3) + sin(]/3)] ~ 600 + 1039

The waveforms and phasors are plotted below. Note how the power into the load is positive for
much of the cycle but negative for a significant part of it (ie. some power flows in and out of the
E. Franklin, B. Blackwell p7/50
ENGN4625/6625 2014 eBrick 1 - 8

load and does no useful work). Despite the total waveform amplitude being 2400 W, the average of
the instantaneous power is only 600 W (it would be 1200 W if the load was resistive only). This is
the same as the real component of the apparent power phasor.

2000 2

1.5 600

Normalised current and voltage


1500
S
Instantaneous power (W)

1
1000 1039
0.5

500 0

-0.5 VL
0
0 10 20 30 40 -1
-500 Power
-1.5
Voltage (normalised) IL
-1000 Current (normalised) -2

Figure 2. Plot of instantaneous voltage, current and power waveforms, and phasor diagram.

2.2.4 Why is complex power S = V × I* ?


We need to use a formalism which depends only on phase difference between V and I; that is the
answer should not be changed if the phases of V and I are both changed by the same amount. (e.g.
both the current and voltage in the “B” phase of a 3 phase circuit are shifted by 120°). This is easy
to see if you simply consider the phase difference between V and I, and expressing V × I* in polar
form, we get |V||I| (∠V-∠I) which is also dependent only on the phase difference between V and I
(Note that the phase angle of a product is the sum of the phases, but the conjugate of I introduces
the minus sign).

Example: Calculate the power when Z = 1Ω (i.e. real), and V = 2+j. (so |V|=√5).

S = V× I* = (2+j)(2-j) = 4 – (–1) + j (2-2) = 5 Watts (Real only: reactive power = 0).

This is correct as V and I are in phase, so we expect P=V2/R, and it is. If we had (incorrectly) used
V× I, (omitting the *) we would have calculated 3Watts + 4[VAr] (4 units of reactive power).

Remember Power = Real(S), and VArs = Imag(S). Power usually means real power, unless we say
“complex power”.

2.2.5 Example: Obtaining Z, given power and power factor


Lagging PF consumption of positive reactive power, care with conj: S=VI*. Lag corresponds to
a negative phase angle (e.g. ∠-60°) for (current) waveforms, but a positive impedance phase angle
(as Z=V/I: current lags in an inductor, phase is negative, but division makes phase or imaginary part
of Z positive).

Given we have a largely inductive (i.e. lagging) load. Z has a positive imaginary part, and as
Z=V/I, V=ZI, so S=VI* = Z(I.I*), so S=Z|I|2. That is, S has the same sign imaginary part as Z.

So i = +h ∗ = oh. h ∗ = o|h| and hence o = i/|h| ( 10)

E. Franklin, B. Blackwell p8/50


ENGN4625/6625 2014 eBrick 1 - 9

i = +h ∗ = +( )∗ = |+| /o ∗ and hence o = |+| /i ∗


p
q
Similarly ( 11)

- very handy for circuit calculations, but be careful of the 1/S*, the sign flips twice (division and
conjugate).

hr
2.2.6 Extension of Power Factor to non-Sinusoidal Waveforms:

4j = × s4j
hr
The non-linear power factor ( 12)

here we consider both the linear power factor treated above (called the Displacement Power Factor
(DPF) and the fraction of the RMS current that is in the fundamental frequency. See Mohan p43,
but we will postpone covering this in more detail until the Power Quality Labs.

2.2.7 Power Factor Correction


Many electrical loads result in current and voltages being out of phase and hence both real and
reactive power must be supplied to the load. The most common reactive load is inductive (eg
motors and transformers) and so we usually have a lagging power factor. As we have discussed, a
power factor less than 1 is undesirable since it places additional burden on the network in terms of
current capacity. Network operators are therefore particularly concerned with loads having a net
power factor as close to unity as possible (in fact so are large customers since, unlike households,
they are actually charged for both real and reactive power). Power factor correction can be at the
customer site, but also network operators employ active power factor correction on their network to
maintain stability and maximise power factor.

The simplest and most common method for power factor correction is by adding reactive
components (eg capacitors or inductors) across the load. Consider the circuit and corresponding
phasor diagram in the figure below, where the capacitor has been added across a load with lagging
power factor. The power factor at the load remains unchanged, but the power factor supplied by the
ac system is now much closer to unity. The same amount of real power is still supplied by the
source but a vastly reduced amount of reactive power is supplied from it. The balance of the
reactive power is supplied by the capacitor.

Figure 3. Power factor correction capacitor added across a load with lagging power factor.

Of course the value of the capacitance required depends upon load itself. And if the load had a
leading power factor then inductance would be required instead. For the power factor correction to
E. Franklin, B. Blackwell p9/50
ENGN4625/6625 201 eBrick 1 - 10
2014

work well the load characteristics must then either be known,


known or otherwise the power factor
correction element must be varied to match the load power factor. This latter method is employed
by utilities, achieved by means of switching capacitors
capacitors and inductors in and out of the power factor
correction circuit (switching is usually done by means of solid-state
solid state devices such as thyristors).

Finally, since non-linear


linear loads such as many power conversion devices can introduce unwanted
harmonics (we willill look at this a little later) into the system as well as resulting in leading or
lagging power factors it is worth looking at an example of combined harmonic filtering and power
factor correction, built-in in this case to a 12 pulse converter. Again, we will look at this converter
later so don’t worry about the details now other than the fact that the converter draws currents from
the supply which consist of the fundamental frequency plus several significant harmonics. The
harmonics are removed by the harmonic
harmonic filters, which at the fundamental frequency also act as
power factor correction capacitors. The remaining capacitors can switched in according to the
reactive power draw at different loads.

Figure 4.. Combined power factor correction and harmonic filtering (from Mohan p470)

E. Franklin, B. Blackwell p10/50


ENGN4625/6625 2014 eBrick 1 - 11

3 Three Phase Power


Power distribution systems are almost universally three phases, spaced uniformly by 120° as shown
in the figures, for practical and economic reasons.

3.1.1 First - Why do we use AC?



1.2
Transformers: change voltage/current level to suit 1
mechanical design considerations 0.8
0.6
– for long-distance transmission, high voltage low current 0.4
2
reduces I R loss, c.f. 0.2
0
– low voltage for safe distribution inside buildings. -0.2 0 5 10 15 20

• Transformers: isolate for safety – no electrical path -0.4


VA
VB
-0.6
between primary and secondary circuits. -0.8
VC

• Fuses, Circuit Breakers interrupt AC much more readily -1


-1.2
than DC – regular current zeros.
• Sliding contacts are much simpler (or not required at all) in AC motors
higher currents and powers are possible.

DC motors must have a multi-segment commutator (complicated, failure-prone)


AC motors can use “slip rings” - simple sliding contacts, or induction, (no contact!).

3.1.2 Why three phase?


• More efficient transmission (less copper)
– regard as three separate circuits, (Vo, Vo∠-120°, Vo∠-240°), return currents cancel
– no return conductor required in theory for balanced circuits
(in practice smaller gauge “neutral” conductors are used at the user end of the network.)
half the conductor compared to single phase!
• Natural source of rotation for machinery (in principle a two phase, quadrature (Vo, Vo∠-90°),

Power is constant during a cycle – +t + +u + +v = Const (= 4152 for domestic three phase).
system would work too, but neutral is not balanced, so still need three wires)

• Interfaces better with rectifiers – load current has much less ripple, line current drawn is
closer to a sinusoid than single or two phase.
• Naturally extends to 6 phase by simple inversion using a 3φ transformer. (invert Voltage)

3.1.3 Three phase Rotation Conventions


Phasors are assumed to rotate anti-clockwise in time as shown in fig 2a. Va (phase a) is taken to be
the reference phase, so that its phase angle is 0 by definition. phase b lags by 120°, so at that
instant, it will be drawn at –120°, or “~7o’clock” (fig. b), and phase c. Note that this requires that
we order the phases CW, so that it is consistent with rotation CCW.

VC –e.g. 240∠-240°
(240° lagging ≡
120° leading)
Va
rotation
Va Va 0°–e.g.
240∠0°
t=0 t =T/3 Va t =2T/3 Vb e.g. 240∠-120°
(120° lagging)

Figure 5 : a VA phasor at 3 consecutive times b/ three 3-phase phasors all at t=0

E. Franklin, B. Blackwell p11/50


ENGN4625/6625 2014 eBrick 1 - 12

3.1.4 Currents and voltages in balanced three phase networks


Note: In the literature, Vphase, Vphase-neutral, Vline-Neutral all mean the phase-neutral voltage (e.g.
VAN , typically 240V), whereas Vline-line refers to the voltage between phases (e.g. VAB, typically
415V). We will try to avoid ambiguity by using VLine-Line (VLL)and VLine-Neutral (VLN), not VLine.

To demonstrate some important features of three phase networks, consider the three separate single
phase circuits of figure 3. Each circuit shares a common neutral potential but has its own return
wire, and identical load impedance. If the three single phase voltage sources have the same
magnitude, V, but have phases of 0°, -120° and 120°C, then:

+& = + = +, +' = + =− +− +, +z = + =− ++ +
Yx Yx
< √X √X
y y

The total current that returns in the three neutral wires is sum of Ia + Ib + Ic:

+& +' +z 1 1 √3 1 √3
h& + h' + hz = + + = {+ − + − +− ++ +| = 0
o o o o 2 2 2 2

Hence the return wires could all be combined into a single wire, which for a balanced set of circuits
carries no current and can therefore even be removed without any impact. The circuit then becomes

load. This is an important result: For a balanced three phase system }~ + }• + }€ = •.


a three-phase circuit with what is known as a Wye connected source and also a Wye connected

Va Vb Vc Ic Ib Ic

Z Z Z

Figure 6. Three separate single phase circuits, can be compared to a Wye connected 3-phase circuit

Another important relationship for three phase systems is the relationship between line-neutral and
line-line voltages. In the circuit shown Va, Vb, and Vc are the line-neutral voltages, and it is
straightforward to calculate the line-line voltages Vab, Vbc ,and Vca:

1 √3 3 √3 ‚
+&' = +& − +' = + − {− + − +| = + + + = √3+ ƒ
2 2 2 2

Similarly +'z = √3+ and +z& = √3+


x „x
Y …

Another important result: in a balanced three phase network line-line voltages are à times
larger in magnitude and are phase shifted +30° compared to line-neutral voltages. This is also
shown in the phasor diagram of figure 4 (the line-line phasors are not drawn from the origin in this
case to help visualise the graphical subtraction of vectors)
E. Franklin, B. Blackwell p12/50
ENGN4625/6625 2014 eBrick 1 - 13

VC Vca

Vbc Va

Vb
Vab = Va - Vb

Figure 7. Line-neutral and line-line phasors shown for a balanced three-phase system

3.1.5 Three phase circuit connections – Wye and Delta


Three phase sources or three phase loads are typically connected up in one of two configurations,
known as Wye (or ‘Y’ or Star) and Delta (or ∆) connections, or a combination of these.
Configurations are shown below, with line-neutral admittances shown for Wye connection and line-
line admittances shown for Delta connection.

C
“Wye” “Delta”
C
YCA
YC
YA
A YBC A
N
YB
YAB
B
B

Figure 8: Wye (or star) and Delta three phase loads

3.2 Balanced and Unbalanced loads


A balanced load is one in which the complex impedance in each arm is the same. A balanced
source is one such as in Fig 2b in which the three source voltage are equal and the phase difference
is 120°. If a balanced load is connected to a balanced source, then by symmetry, the voltage at the
neutral node “N” will be zero, regardless of whether it is connected or open circuit. Later, the
symmetric component transform will let us treat balanced circuits as a single circuit, and by
decomposing, simplify the treatment of unbalanced circuits.

It is convenient to introduce a 120° rotation operator, ℎ = − +


3.2.1 The h operator and the H matrix
√X
- h is a complex cube root
of 1, the other cube roots are ℎ (= ℎ ) -. ℎX (= 1). The operator can also be written in polar co-

ordinates, ℎ =
Yx
y .

So for a balanced three phase set of voltages (sequence ABC) we can write [+t , +u , +v ] =
+t [1, ℎ , ℎ]. The H matrix will be introduced later in the symmetric component transform.
E. Franklin, B. Blackwell p13/50
ENGN4625/6625 2014 eBrick 1 - 14

3.2.2 Example Calculation with neutral intact – by inspection.


Consider an unbalanced load: A solidly grounded star connected load comprises impedances of:

Za = 10 Ω ∠15° , Zb=12Ω ∠7° & Zc= 4Ω ∠32°

C 3 phase rotation operator h := − 1/2+ j


√3
“Wye” 2
VC
ZC 15 π
ZA j
180 V A := 240
N A Z A := 10 e = 9.659+ 2.588 i
IA j

2
Z B := 12 e 180
= 11.91+ 1.462i V B := 240 h
ZB IN
VA 32π
j
Z C := 4 e 180
= 3.392+ 2.120 i V C := 240 h

B
VB If the supply and cables have negligible impedance, and the neutral
is well grounded,
I A := V A /Z A I A = 23.18+ -6.212i ∣I A∣= 24.00
I B := V B / Z B I B = -12.04+ -15.97i ∣I B∣= 20.00
I C := V C /Z C I C = 2.094+ 59.96 i ∣I C∣= 60.00
I N := − ( I A+ I B+ I C ) I N = -13.24+ -37.78 i ∣I N ∣= 40.03
Figure 9: Unbalanced star load with neutral intact

This shows a large neutral current, a result of the imbalance in magnitude and the deviations in
phase of the currents in the three unbalanced loads.

3.2.3 Three wire connection (no neutral)


Sometimes there is no neutral connection, either by design or by accident. Then the circuit
equations cannot be solved so easily by inspection and all three currents interact. We will introduce
the nodal admittance matrix method in the next section. In Power systems, the sources are almost
always voltage sources, with low impedance, and the loads are usually in parallel. Therefore it
makes a lot more sense to use admittance, since for parallel loads the admittances (and also the
currents) simply add.

3.2.4 Power in polyphase (eg. three phase) circuits


In polyphase circuits (three phase being of course the most common), for the phase voltage vector
V and phase current vector I, the general expression for total system complex power is given by:

i = + L × h∗ ( 13)

i = +& h& ∗ + +' h' ∗ + +z hz ∗


The transpose is necessary to keep the matrix maths consistent. For a three phase system, the total
complex power is thus simply: ( 14)

In the above expressions, V may actually either be the line-neutral voltages and line-neutral currents
(eg for a Wye connected load), or may be line-line voltages and line-line currents (eg for a Delta
connected load). Since the voltage and currents in the loads are different by a factor of √3 in an
ideal balanced circuit (if balanced VCA = √3 VCN), it therefore very important to be consistent
E. Franklin, B. Blackwell p14/50
ENGN4625/6625 2014 eBrick 1 - 15

Note: if we measure for example VCN, then we should use the corresponding current into terminal C
(IC) when calculating power. VCN is known as VLINE-NEUTRAL or VLN (sometimes referred to as
VPHASE). IC is known as ILINE-NEUTRAL or ILN (also sometimes referred to as ILINE). Similarly VCA is
known as VLINE-LINE, or VLL, and ICA is known as ILINE-LINE, or ILL. To avoid confusion, try to
always use the long form in each case. Also note that IC = ICA+ICB or ICA-IBC

The real power in a balanced system is given by: 4 = 3 +‰Š h‰Š cos = 3 +‰‰ h‰‰ cos ( 15)

4 = √3 +‰‰ h‰Š cos = 3√3 +‰Š h‰‰ cos , which is confusing and hence best to avoid!
However if we were to use mixed quantities, then our expression would become equivalent to:

4 The Nodal Admittance Matrix method


Power networks are characterised by having a large number of nodes whose voltage is referenced to
a common node. Since voltage sources are referenced to this common node and the number of
voltage sources is usually small in relation to the number of nodes, the Nodal Admittance form of
solution is highly suitable. This form of solution has the additional benefit that the number of
equations needed to solve the network is at all times equal to the total number of nodes minus one
(which is the ref. node). In contrast, if a mesh impedance solution is attempted, the number of
independent equations depends on the number of nodes and branches i.e. it depends on the precise
topology.

The Nodal Admittance Matrix method solves for the injected current from voltage generators i.e.

I=Y· V

I is the n x 1 matrix (or column vector) representing the currents injected into each node, V is n x 1
matrix representing the voltage at each node of the network, and Y is an n x n matrix, the so-called
Nodal Admittance Matrix.

For large networks the admittance matrix Y will of course be very large (and as you will see is
usually a sparse matrix – lots of zeros), however by using this method the solution to any network
can be reduced to a series of matrix manipulations. These are time-consuming (to say the least) by
hand, but can be handled very reliably and efficiently by computers. This is the basis for all
complex power network solutions. The method also forms the basis for computer simulations such
as Spice, which solve this equation at each discrete time-step (with the added complication that
iterative solutions are required when non-linear elements such as diodes are included).

4.1 Determining the NAM by inspection


Consider the section of network shown in the figure. It has
2
three nodes (1, 2 and 3) and a common reference node
I2
(node 0). Voltage sources are shown to make it clearer, but Y12
the essential thing to note is that the voltage at each node i
is Vi relative to the common reference node and the injected 1
I1
current into that node is Ii.
Y13 3
V2 V1 Y10
In the column vector I non-zeros only occur at nodes where
a generator is connected (ie the sum of currents into the O 0 is common
node must equal zero if there is no external source). The
E. Franklin, B. Blackwell p15/50
ENGN4625/6625 2014 eBrick 1 - 16

entries in Y can be easily written down by inspection of the network.

Starting with node 1 which has a generator connected, a shunt admittance to the common node 0
and connections to several other nodes. Applying Kirchoff’s Current Law to express injected
current I1 in terms of voltages and admittances:

Current into the node Current flow to other nodes To node


I1 = Y10 *(V1 – V0) = Y10V1 – Y10V0 0
+ Y12 *(V1 – V2) = Y12V1 – Y12V2 2
+ Y13 *(V1 – V3) = Y13V1 – Y13V3 3
+…………. …
+ Y1N *(V1 – VN) = Y1NV1 – Y1NVN N

I2 = Y20 *(V2 – V0) = Y20V2 – Y20V0 0


…… + Y21 *(V2 – V1) = Y21V2 – Y21V1 1
…… + Y23 *(V2 – V3) = Y23V2 – Y23V3 3

Re-arranging these terms gives

I1 = (Y10 +Y12 + Y13 +….. +Y1n)*V1 + (-Y12*V2 –Y13*V3 - …… - Y1n*Vn)

I2 = (Y20 +Y21 + Y23 +….. +Y2n)*V2 + (-Y21*V1 –Y23*V3 - …… - Y2n*Vn)

etc. where the colours show where the terms came from

This leads to the general rules for the formulation by inspection of the Nodal Admittance Matrix for
any power network:

1. The matrix Y is symmetrical


2. The self-admittance Yii (each of the diagonal terms) is equal to the sum of all admittances
connected to the ith node (red text above)
3. The mutual admittance Yij (each of the off-diagonal terms) is equal to the negative of the
admittance between the respective nodes i and j (blue text)

Thus the nodal admittance matrix can be written down in an automated manner without any
reference to network topology apart from noting which nodes each admittance connects to. Here it
is written for the previous network segment but extended for a general N node network.

S < + S + ⋯ S Š −S −S X …
−S S < + S + ⋯ S Š −S X …
S=‹ …•
−S X −S X SX< + SX + ⋯ SXŠ

… … …

E. Franklin, B. Blackwell p16/50


ENGN4625/6625 2014 eBrick 1 - 17

4.1.1 Example Calculation Open neutral: Z A= 10 Ω∣ 15 , Z B= 12 Ω∣ 7o , Z C = 4 Ω∣ 32o


o

The problem is the same as before but 3 phase rotation operator h := − 1/ 2+ j


√3
2
the neutral is broken. 15 π
j
Z A := 10 e 180
= 9.659+ 2.588 i V A := 240

j 2
C Z B := 12 e 180
= 11.91+ 1.462i V B := 240 h
“Wye”
VC
32π
j
ZC Z C := 4 e 180
= 3.392+ 2.120 i V C := 240 h
ZA
N A
IA
IN = 0
ZB Broken VA In this case, we will use the Nodal Admittance Matrix method, so
connection
get Y's from Z's
B
VB Y A := 1 /Z A Y B := 1/Z B Y C := 1/ Z C

If the supply and cables have negligible impedance, and the neutral
is open circuit, we obtain Y by inspection (rules in eBrick 2)
Figure 10: Unbalanced load with broken neutral

[ ]
YA 0 0 −Y A
The nodal admittance matrix is very 0 YB 0 −YB
Y :=
simple to write in this case, many 0 0 YC − YC
− Y A − Y B − Y C Y A+ Y B + Y C
terms are zero – e.g. no connection
across phases. We need to evaluate I := Y V , but first we need V N
I N = − ( I A+ I B+ I C )= 0 , or from the bottom line of the matrix,
The trick is that we don’t know VN, V N := (Y A V A+ Y B V B+ Y C V C )/(Y A+ Y B + Y C )= -6.530+ 93.73 i
so we have to solve for it. If we look
We see the neutral voltage is far from zero!
back at how the nodal admittance was V := [V A ,V B , V C ,V N ]
Now make the vector V:
obtained in the section above, at is

[
21.39+ -15.43i
essentially setting the bottom line of
the equations to zero on the LHS
(IN=0). Or you could just use KCL.
then I := Y V =
-12.45+ -23.79i
T

-8.939+ 39.23i
0
]
and ∣I 0∣= 26.37 ∣I 1∣= 26.85 ∣I 2∣
= 40.23 ∣I 3∣= 0
We see also that the neutral current is zero as expected.

The technique we used in the above example essentially to remove the neutral node N can also be
generalised to remove any number of internal nodes. This is done via a technique known as matrix
partitioning. The matrices are first reordered such that the matrix equation can be re-written as:

h ” • +•
h = S. + → • • ’ = “ —• ’
h‘ – s +‘

Here A, B, C and D are different size sub-matrices chosen in such a way that Is corresponds to all
the source currents (which we are trying to find solution for), Ii are the internal nodes with no
current injection (=0), Vi are the internal node voltages (which we don’t know values for) and Vs are
the source voltages.

The matrix can be expanded out: IS = AVS + BVi


Ii = CVS + DVi
And hence for Ii = 0, Vi = -D-1CVS

And therefore IS = AVS + B(-D-1CVS) IS = (A - BD-1C)VS ( 16)


E. Franklin, B. Blackwell p17/50
ENGN4625/6625 2014 eBrick 1 - 18

The admittance matrix equation (containing unknown node voltages) has thus been reduced to a
simple calculation matrix calculation with known source voltages only. After the solution has been
found, internal node voltages can then easily be calculated if required.

Going back to our prevuious example, the NAM can be re-written in the same format, with:

Y A 0 0 − YA 
A =  0 YB 0  , B = − Y  , C = [-Y –Y –Y ] and
 B A B C D = [YA+YB+YC]
 0 0 YC  − YC 

It can be easily shown that solution for Is becomes:

œ St +t − St D •ž •Ÿ • F£
•ž pž •Ÿ pŸ • p

ht +t
› ¢
hr = ˜hu ™ = (” − •s –) ˜+u ™ = ›Su +u − Su D • • • F¢
•ž pž •Ÿ pŸ • p

hv +v › ¢
ž Ÿ

š Sv +v − Sv D •ž •Ÿ • F¡
•ž pž •Ÿ pŸ • p

Which is of course the same result as we obtained previously.

4.2 Star-Delta transformation via the NAM


Consider two of the simplest three phase networks below, called the wye (“Y”) or star, and delta
networks. In our course, neither the Y nor ∆ are in the usual orientation, because we try to draw the
components so that their terminals coincide with the corresponding phasor voltages.

C
“Wye” “Delta”
C
YCA
YC
YA
A YBC A
N
YB
YAB
B
B

Consider the nodal admittance matrices for Wye (Y) and Delta (∆) networks. We’ve already looked
at the Wye network; writing the matrix for the Delta network is even simpler:

St 0 0 −St
Stu + Svt −Stu −Svt
0 Su 0 −Su
S• = ‹ 0 Sv •, S∆ = ˜ −Stu Stu + Suv −Suv ™
0 −Sv
−St −Sv −Svt −Suv Svt + Suv
−Su St + Su + Sv

E. Franklin, B. Blackwell p18/50


ENGN4625/6625 2014 eBrick 1 - 19

Question: What is the equivalent “Delta” circuit to a general “Wye” or “Star” arrangement of
admittances YA, YB, YC? – i.e. what component values make the two circuits above equivalent
from the point of view of their external nodes A, B and C?

Call the Delta equivalents YAB, YBC, YCA, as they connect two nodes, in contrast to the Wye with a
single subscript, as admittances here go from one node to neutral. We could obtain the result from
general nodal analysis, however it is even simpler to show, by comparing the Nodal Admittance
Matrix for both circuits, that

St Su )
Stu =
St + Su + Sv
Star→Delta ( 17)
This is referred to “straddle” / “sum”. (meaning product of the straddling nodes over the sum of all
nodes). For the special case YA = YB = YC = Y, YAB = Y/3. etc. In other words, a star
arrangement of 10Ω loads would be indistinguishable from a delta arrangement of 30Ω loads (in
terms of any measured external currents or voltages).

Conversely, the delta to star transform is of the same form, but in terms of Z’s
e.g.
otu ovt
ot =
Delta→Star otu + ouv + ovt ( 18)

If expressed in terms of Y’s, this looks a bit different: 7 Y terms, and is harder to remember.

For balanced three phase loads, Wye and Delta connections can be considered interchangeable by
the rest of the network provided that Y∆ = YY / 3, or Z∆ = 3.ZY ( 19)

Using this transformation makes calculations for a combination of star and delta loads trivial to
incorporate: e.g. the delta connected 10kVA load in Tutorial questions 1 - transform to Star (YY =
3Y∆ ), then add in the admittance to the first three (single phase) loads.

5 Power Transmission
We have seen already some of the reasons why electrical power systems normally use 3 phase
power. Power networks typically consist of combinations of generators, transformers, and
transmission and distribution lines, and distributed loads. In the circuits we have considered so far
we ignore any effects of line impedance. In practice however, voltage sources (generators or
transformers) and loads are typically spaced apart by some distance and the transmission lines
joining them have some non-zero impedance associated with them which must be taken into
consideration for real power networks.

Another (usually reasonable) assumption that we make is that voltage sources maintain a constant
supply voltage regardless of load. In reality voltages are also supplied by real transformers which
usually don’t maintain constant voltage as load is increased, although network operators can use

E. Franklin, B. Blackwell p19/50


ENGN4625/6625 2014 eBrick 1 - 20

voltage regulation at the transformer to combat this. This is dealt with in more detail later in the
transformers section of the course.

As an interesting aside, in Australia the low voltage distribution network operators are required to
deliver mains voltage within the range 230V (or 398V line-line for 3-phase) +10% -6%. The
standard used to be 240V (415V for 3-phase) the majority of distribution networks will still operate
at this voltage until hardware is gradually changed. The network operator will check the local
network carefully before allowing a retail customer to connect a large new load to the network, or a
large source (such as a large PV system), since this can easily bring the mains voltage out of spec.

5.1 Impedances of transmission lines


Transmisson line cables or conductors are composed of multiple strands of aluminium wire wound
together (often on a steel stranded core for strength). We are already familiar with the concept of
the linear resistance of a piece of wire (which of course is dependent upon the metal resistivity, the
cross-section over which current flows and the length of the wire. However there are three other
common impedances to consider in real transmission systems. These are associated with:

- series resistance of conductors


- self-inductance of each conductor and linked inductance between conductors
- cable to ground or cable to cable shunt resistance (very high but not zero),
- cable to ground or cable to cable to capacitance

While conductor series and shunt resistance can be an


important consideration for some transmission line
analysis, generally speaking their impact is considerably
D D less than that of inductive and capacitive effects.
Consider the cross-section of a three-phase transmission
line shown in the figure. In this case the three
conductors, each carrying a separate phase, are spaced
D equally apart a distance D. Each conductor has a radius
R.

Each conductor experiences an induced voltage due to magnetic flux generated by itself and also
from both of the other two conductors. Likewise the capacitance associated with one conductor is
influenced by the charge on each other conductor. However, in the normal case where the system is
balanced (ie each conductor carrying same current with normal phase sequence) then the per unit
length line inductance and line to neutral capacitance can be given by the simple expressions
(derived in Bergen Chapter 3):

¦= ¦-
§¨ ©
, where -< is the permeability in free space
‚ <.ª« ¬
H/m ( 20)

®=
2]¯0
s , where ¯< is the permittivity in free space
¦-(
F/m ( 21)

It can be seen that inductance can be reduced and capacitance increased by decreasing the spacing
D or by increasing the conductor radius. Usually inductance is more significant and so such a
change is desirable. However D can only be safely reduced so far; the solution usually employed is
E. Franklin, B. Blackwell p20/50
ENGN4625/6625 2014 eBrick 1 - 21

to bundle conductors (the characteristic close-spaced square or triangular layout of conductors that
is often seen in transmission lines) which effectively increases the conductor radius. In this case D
is replaced by a quantity known as the geometric mean distance (GMD) between bundles and R is
replaced by the so-called geometric mean radius (GMR) of the conductor bundles.

5.2 Transmission line models


For a three phase transmission line operating under balanced conditions, a per phase lumped model
for a transmission line can then be represented by the figure below.

sinh ³¦
o ° = ±¦
³¦
I1 Z' I2

³¦
S ° N¦ tanh 2
V1 V2

=
2 2 ³¦
Y'/2 Y'/2

³ = µN±

Figure 11: A per-phase transmission line model

Here z is per unit length line impedance (z = r + jwl) and y is the per unit line-neutral admittance (y

lossless line case where r = 0 & g = 0, the propagation constant ³ = _√¦®.


= g + jwc ). Transmission line length is l. For the case where wc >> g and wl >> r , or for the

¾¿
¶·¸¹ º» ¼½¸¹
Y
º» ¾¿
- For ‘medium length’ transmission lines (less than about 240 km), both and are
Y
typically close to 1. Hence the same lumped model can be used but with much simplified
parameters:

o ° = o = ±¦ and S ° = S = N¦

The so-called characteristic impedance Z0 of the transmission line is given by o< = CÁ = C• which
À q

for the lossless line becomes • = CÄ


Ã

- For ‘short’ transmission lines (less than about 80 km) the admittance can be neglected
entirely and the lumped model is reduced to a simple line impedance Z = R + jwL

This brings us to a general model that we can use for most three phase networks where line
impedance needs to be considered. A simple three phase system with transmission line impedances
included is shown in figure 9. Recalling that unbalanced currents result in mutual inductances M
between phases, the transmission network for a more general unbalanced network is also shown.

E. Franklin, B. Blackwell p21/50


ENGN4625/6625 2014 eBrick 1 - 22

R L

M
R L M

M
R L

Figure 12: A three network with transmission line impedances (top) and mutual inductances, M,
between phases shown. Note that for a balanced network, mutual inductances are not present.

5.3 Propagation on transmission lines


A full consideration of a transmission line, taking into account the variation in voltages and currents
with position along the length of the line owing to the per unit length impedance and admittance,
reveals that voltage and current signals propagate along the line in a finite time period. Detailed
analysis of the travelling waves on transmission lines is not covered in this course, but we will note
a few important properties. Firstly, the voltage and current waves propagate with a characteristic
velocity of propagation, v:

Å=
Æ
√Ç€
( 22)

section) the velocity is Å =


Æ
In the limiting case of conductor radius R << spacing between conductors D (from previous

√Ç€
= 3 x 108 m/s (ie. the speed of light). In practice the speed of

speed of light). The wavelength È = of a propagating 50Hz signal is thus (for v = 66% of speed
Å
propagation is somewhat less than this, but is still nonetheless fast! (typically 60% to 80% of the
É

of light) 4000 km. And the propagation delay Ê = Å for a 40km line (for eg) would be 0.2 mSec.
Ç

È=É
Å
Wavelength of oscillating signal on transmission line:

Ê=Å
Ç
Propagation delay for a transmission line length l:

VS
ZT
Transmission line, Z0

Figure 13: A transmission line shown with termination impedance ZT

A second important property is that propagating waves reflect at the end of a transmission line and
a reflected wave propagates in the reverse direction. The reflected fraction is given by the following
reflection coefficients Rv and Ri for voltage and for current respectively, where Z0 is the
E. Franklin, B. Blackwell p22/50
ENGN4625/6625 2014 eBrick 1 - 23

characteristic line impedance and ZT is the termination impedance at the terminated end of the line
(eg a load impedance). The same applies at the source end of the line, with the source impedance

(Ë = (‘ = −
qÌ q¨ qÌ q¨
qÌ q¨ qÌ q¨
and ( 23) & ( 24)

There are three special termination cases:

1. If the line termination is open-circuit, then Rv = 1 and Ri = -1. The voltage at the open-
circuit terminal will then be twice the ‘sent’ voltage and the current will be zero. This is a
particular concern for designers when considering voltage surges (a lightning strike an
extreme but real example).
2. If the line termination is short-circuit, then Rv = -1 and Ri = 1. The voltage at the open-
circuit terminal will then be zero and the current will be double.
3. If the line termination ZT is matched to the characteristic impedance Z0 , then Rv = 0 and Ri
= 0. There is then no reflected wave. This is especially important for communication
systems, audio systems (you are probably familiar with this already)

The two travelling waves (one in each direction) interfere constructively or destructively, and for
lossless lines (no resistance elements) of certain length (eg multiples of quarter wavelengths – 1000
km for our previous example) and with particular termination conditions standing waves can even
be set up. These can be handled the same way and have similar consequences as for waves in other
mediums (eg sound waves). In all other situations the reflected waves are of diminishing magnitude
after each reflection and hence die out after a given time period, resulting in classic damped
oscillation (see figure below).
40KM Line example; Vin = step of 1000. V magnitude.
Voltage into line and at 20km and open circuit end.
2000
[V]

Open circuit 40km power


1600
transmission line: Voltage
overshoots by 70%
1200

800

Source = red
400
green - mid (20km)
blue - end (40km)
0
0 1 2 3 4 [ms] 5
(file anusurge01.pl4; x-var t) v:SURGE v:20KM v:40KM

Figure 14: Propagation of a constant voltage applied to source end of a 40km transmission line with
propagation delay equal to 0.22 ms.

The frequency of the oscillation (or ringing) is determined by the propagation time of the line and
the whether the reflected signal is inverted at one or both ends. The signal decay in time (damping)
is determined by the magnitude of reflection coefficients and the propagation delay. For a lossy line
(series resistance and shunt conductance not negligible) the travelling wave is also attenuated
during propagation.

E. Franklin, B. Blackwell p23/50


ENGN4625/6625 2014 eBrick 1 - 24

TÍ•z =
ÎÏ
If signal is inverted at one end and not other, we have two round trips per cycle:

TÍ•z =
Ï
If signal is inverted at both ends, we have one round trip per cycle:

6 Coordinate Transformations
6.1 Introduction
Just as transformation to polar coordinates can simplify analysis, transformation of circuit
quantities to different basis sets is useful. The concept is to change to a frame where the variables
are decoupled from each other, ideally producing equations in one or two variables instead of three,
six or more. A simple example is the floating neutral example, which caused the currents and
voltages to be coupled, compared to the earlier grounded neutral case, where each phase could be
analysed separately. Even with grounded neutrals, mutual inductance between windings in motors
act to couple the circuit of one phase to another. Rotating machinery is particularly complicated as
the inductance values vary as the rotor rotates.

We will consider two such transformations, the Symmetric Component Transform which can
simplify any three phase system, and the DQ0 transform which works in a frame that rotates with
the shaft of a rotating machine (motor or generator). The latter works because it transforms the
variation of rotor and stator inductance with angle into just another variation with time.

6.2 Power Invariance


Ideally transformations would be power invariant – the power calculated in one frame would agree
with the power in another. Consider two systems – ABC (normal 3 phase) and 012. Writing
Ohm’s law in our familiar system:

+tuv = otuv htuv ( 25)


Where the impedance matrix is of the form
ott otu otv
otuv = out ouu ovv
ovt ovu ovv
( 26)

section) is also 3x3 – 3 coefficients for each component of the original quantity (e.g+t , +u , +v . ) to
A general transformation matrix (for example the symmetric component or H matrix in the next

make the 3 new quantities. Take such a transformation matrix A (from 012 to ABC)

Let +tuv = ” +< and htuv = ” h<

Substitute in ( 25) and multiply on left by A-1.

+< = ” otuv ” h< = o< h< ( 27)

hence o< = ” otuv ” ( 28)

Next calculate the power


∗% ∗%
Power (3 ph) Re +tuv
L
htuv

= Re$ ” +< L
”h< = Re$ +< ” ”∗ h<
L L
( 29)
E. Franklin, B. Blackwell p24/50
ENGN4625/6625 2014 eBrick 1 - 25

So power is conserved iff ÐÑ Ð∗ = }Ò, the identity matrix.

(The real variable counterpart is ÐÑ Ð = }Ò e.g. DQ0 transform below)

This is an important property – if it doesn’t hold, then you have to transform back to the original
frame to calculate power. The following particular transform, the Symmetric Component
Transform, doesn’t have the power invariance property.

6.3 The Symmetric Component Transform


The Symmetric Component Transform (SCT) is a powerful transformation tool which allows us to
represent unbalanced networks in terms of balanced networks only. We will see how it also allows
for simplification of network analysis considerably, and even for balanced networks it can greatly
simplify the equations governing the network.

The basis of the symmetric component transform is the theory that any arbitrary set of three phasors
(for example the three phases of an unbalanced 3-phase system) can be represented instead by three
balanced ‘sequence sets’ of three phasors (ie 9 phasors in total). Specifically the three sequence sets
are known as the positive-sequence set (phase sequence ABC), the negative-sequence set (phase
sequence ACB) and the zero-sequence set (all in phase).

These sequence sets were introduced by Fortesque (in 1918) who noticed that for static components
(not rotating like motors and generators), the diagonal (‘self’ impedances) of the impedance matrix
such as ( 26) has values almost the same, and the off-diagonal (‘mutual’ impedances – see box at
top) are almost the same as each other.

o Self impedances are very similar


o Mutual impedances are very similar (but different to self)

The symmetric component transform notation we use is that the positive sequence set is represented
by V1 (and I1), the negative sequence set is represented by V2 (and I2), and the zero sequence set is
represented by V0 (and I0). Note that some texts use V+, V-, V0 for these sets. We should note

for the positive sequence set we can denote the three voltage phasors as +& , +' and +z . Of course
further that each set consists of three phasors which we can still denote as a,b, and c; for example

since each set is a balanced set, it is possible to relate the phasors to one another, ie , +' = ℎ +&
and +z = ℎ+&

The transformation makes each of +t , +u , +v out of a linear combination of +< , + , + and similarly
for current I. Working in currents (the same can be done though for voltages), we can easily write
the original three current phasors in terms of the new symmetric components:

I a  I a  I a  I a 
0 1 2

     
I abc =  I b  =  I b0  +  I b1  +  I b2 
 I c   I c0   I c1   I c2 

E. Franklin, B. Blackwell p25/50


ENGN4625/6625 2014 eBrick 1 - 26

I a  1 1  1  1 1 1   I a 
0

 
which we
I abc =  I b  = I a0 1 + I a1 h 2  + I a2  h  = 1 h 2 h   I a1  = HI 012 ( 30)
expand as
2  2
 I c  1  h  h  1 h h   I a 
2

This takes the three phasors h< , h , h , and multiplies each by three coefficients (1,1,1) for the top
row, which in total gives IA, and similarly for IB 1, ℎ , ℎ and IC 1, ℎ, ℎ , where h is the 120°

a matrix (H) multiplied by the vector h< , h , h . The notation h&< is to remind us that the phasor I0 is
rotation operator (the cube root of 1) as introduced earlier. The combined operation is equivalent to

the same as the “A” phase component of itself when transformed into ABC space. (the top row of H
is all 1’s). The symmetric component phasors are shown diagrammatically below. It must be noted
however that the symmetric components are all shown here to have their A phases in phase with
each other. This is certainly not necessarily the case (as we will soon show with an example).

[I0A, I0B, I0C] , [I1A, I1B, I1C ] and [I2A, I2B, I2C]
IC IB

All in phase
IC rotation rotation
IA 0° e.g. IA 0° e.g.
IB
240∠0° 240∠0°
IA
I0: Zero I1: Positive I2: Negative
Sequence IB Sequence IC Sequence

1 1 1 1 1 1
Ó = ˜1 ℎ ℎ ™ from which it follows that Ó ∗ = ˜1 ℎ ℎ ™
1 ℎ ℎ 1 ℎ ℎ
So ( 31)
1 ∗
Ó = Ó
3
And you can confirm
( 32)

That is to get our symmetrical components from the original ‘real’ phase currents or voltages we
need to use:

 I a0  1 1 1 I a 
 1
=  I a  = H −1 I abc = 1 h h 2   I b 
1
I 012 ( 33)
3
 I a2  1 h 2 h   I c 
 

This tells it that the symmetric component transform is NOT power invariant. This definition is
chosen so that the voltages and currents can be related easily. If we were to redefine the SCT
transform to include 1/√3 in both directions, then the voltages and currents would not be directly
comparable – a [Va,Vb,Vc] set of 11kV each would correspond to a [V0,V1,V2] set of [0,19kV,0],
which would be even more confusing. As above, we just need to convert back to ABC to get the
true power. Because of the simple relationship between components, we usually consider only the

stage by rotation by h. Thus we often write I0 or I0 when strictly we mean h&< .


“a” phase components (called the “lead” or “reference” phase), and obtain the b and c at the last

E. Franklin, B. Blackwell p26/50


ENGN4625/6625 2014 eBrick 1 - 27

6.3.1 Example calculation of symmetric components


Consider an unbalanced three-phase system which has the following voltages (note: these voltages
are per unit voltages – we will cover in some more detail later):

Va = 0.9 ∠0°, Vb = 1.25 ∠280° , Vc = 0.6 ∠110°

We can write the leading symmetric component phasors:

+&< = X 0.9 ∠0° + 1.25 ∠280° + 0.6 ∠110° = X 0.9 + 0.22 − j1.23 − 0.21 + j0.56 = 0.38 ∠ − 36°

+& = 0.9 ∠0° + 1.25 ∠40° + 0.6 ∠350° = 0.9 + 0.95 + j0.8 + 0.59 − j0.10 = 0.85 ∠16°
X X

+& = X 0.9 ∠0° + 1.25 ∠160° + 0.6 ∠230° = X 0.9 − 1.17 + j0.43 − 0.39 − j0.46 = 0.22 ∠183°

The remaining symmetric component phasors can of course be calculated from these by angular
rotation. You can always check that you have got this right by summing to get Va.

It should be quite obvious, but if our three-phase system is balanced and has positive sequence then
if we calculate the symmetric components we would find that +&< = 0, +& = +& and +& = 0. This is
very useful since we can often perform the transformation on impedance matrices and solve
circuits more simply in symmetric components (and the positive sequence symmetric component is
the only one that requires solving).

6.3.2 Simplification of a general, balanced static impedance


As stated previously, a balanced, static (non-rotating) circuit impedance has the following
symmetry
o• oØ oØ
otuv = oØ o• oØ
oØ oØ o•
( 34)

where Zs and Zm are the self and mutual impedances respectively. The symmetric component
transform diagonalises this (try it!).

o• + 2oØ 0 0
o< = Ó otuv Ó = ˜ 0 o• − oØ 0 ™
0 0 o• − oØ
from ( 27) ( 35)

Example: ZABC self (diagonal) impedance = 10 + j25, mutual impedance 0+j7 Ω →

16 + 39 0 0
o012 = Ó−1 o”•– Ó = Ù 0 7 + 18 0 Ú
0 0 7 + 18

The significance of this ( 35) is that we have reduced an arbitrary balanced (static) circuit to three

sequence set, V1), so we are down to one equation: ÛÆ = ÂÆÆ }Æ = ÂÜ − ÂÝ }Æ , other I’s are 0.
simple equations – and if the drive term is balanced, then we only have one source (the positive

Further development of this idea leads to great simplification analysing in the consequences of
faults (short circuits) in power distribution system. The separation into separate circuits for each
component, and some simple properties allow reduction of the network to a much smaller
equivalent. The MEng course will cover this in some more detail.
E. Franklin, B. Blackwell p27/50
ENGN4625/6625 2014 eBrick 1 - 28

6.3.3 Simplification of a general, balanced AC machine impedance


The SCT also simplifies a generalised rotating machine’s impedance:
i Þ U
otuv = ˜ U i Þ™
(we omit the Z,as m and n look too similar
Þ U i
in subscripts)
( 36)

i+Þ+U 0 0
o_012 = ˜ 0 i + ℎ Þ + ℎU 0 ™
0 0 i + ℎ Þ + ℎ U
Transforms to ( 37)

which shows that the impedance of a rotating machine is quite different for different excitations –
in particular the if all windings are connected to the same phase, the current drawn will be very
different to the current when connected to a normal rotating three phase supply.

6.3.4 Important Properties of the Symmetric Component Transform


1. If the circuit is linear, superposition applies, and if the (symmetric) circuit only has sources
just of one sequence(+,-,0), then the responses are of the same sequence.
2. + and – sequence sets are balanced (no neutral current) – so for balanced networks and
loads, you can assume all neutral currents are zero for +,- sequ. (and usually that neutral
voltages are zero).
3. For the 0 sequence (all in phase), any star networks with no neutral connection can be
ignored altogether.
4. Unbalanced faults can be converted to balanced Symmetric Component current sources.
See Line Fault analysis below)
5. Unbalanced sources, when represented in SCT are balanced individually (i.e. for the 0,1&2
cpts) – just the sum will be unbalanced.
6. The “a” phase (lead phase) can be reconstructed from the 0,1,2 SCT components by simple
addition, but don’t forget
– the phase rotation terms (“h2”s and “h1”s) when reconstructing the B and C phases,
– or the 1/3 when going from A,B,C, to 0,1,2.

6.3.5 Applications of the Symmetric Component Transform


General simplification of circuits and or sources: applying especially to faults and motors
Fault analysis:
– Allows use of single line or per phase methods even if the fault is not balanced (most
cases!)
o Not to save computer time! but to allow a better understanding;
o Need to get the right answer of course, but most importantly, develop an
engineering judgement as to the validity of your result.
– Protection: Commonly used in 3 phase protection “relays” (ch 8) as a criterion to “trip”
(disconnect circuit) e.g.
o Detect imbalance in sources – failure of one phase (this could be a neg. seq. cpt.>5%)
o Detect imbalance in load – may be a short
Inferior alternative is to set limits on currents in all three phases, but
Needs three set points
What about starting currents? - May be large, but if the correct sequence, OK
o or a less severe case - faulty motor

E. Franklin, B. Blackwell p28/50


ENGN4625/6625 2014 eBrick 1 - 29

An internal short in a winding simply increases dissipation and heating - current


may be well within limits, power same, but more heat, less output
imbalance will generate a negative sequence component.
Rotating machinery (motors, generators) in general best modelled in this coordinate system
The main parameters (mutual and self impedances) vary greatly with the rotation direction,
so the most natural coordinate system
Example: The positive sequence component of the reactance of a synchronous motor
changes markedly with time after a load change (transient). The neg. and zero don’t.

6.3.6 Negative Sequence Detection in Motors


SCT can be used to demonstrate the importance of negative sequence detection in motors. We can
show this by way of an example:

If we assume a motor can be represented by a delta equivalent circuit, with balanced winding
impedances / admittances and driven from a normal balanced source.
In good condition, with all windings balanced Yab=Ybc=Yca = 10∠ 40°
– SCT of I = [0, 7200, 0] Good – all positive sequence
Assume Yca has a partial short – Y doesn’t change, gets a little more resistive
Yca = 10∠ 30°
– SCT of I = [0, 7176, 418] 7% of power in neg seq, fights pos
– output is only 84% (rest is heat! – burns out?)
So simply detecting the negative sequence component > few percent detects many types of faults as
below
– Missing Phase (i.e. Vabc=240*[0, h^2, h], A phase is zero
– Short in any of the phases
– Subtle change in impedance/admittance as per the above example
Negative sequence relays are used in real systems to avoid damage to motors. The negative
sequence is far more readily detected for unbalanced operation than when any over-current
protection would kick in.

6.3.7 Line Fault Analysis


Line faults occur when a distribution mains conductor (415V and higher) contacts ground (or a
tree). The most common case is just one line shorting, and is called a Single Line to Ground (SLG)
fault. In spite of the inherently unbalanced nature of this circuit, it can be treated very effectively
by Symmetric Component Analysis.

Usually problems with supply imbalance are local to a building or a single circuit. The mains feed
and distribution feed a large number of loads, so the imbalance is averaged in the distribution
circuits where the fault is likely to occur. If it is not, then the supply authority or plant management
will probably do something to correct it, such as rearranging loads into different circuits.

So we can assume a balanced source and a balanced load, apart from the fault itself, which is
invariably highly unbalanced. You can easily show (from ( 32): H-1 = 1/3 H*) that a single line (A
phase) current of 1 Amp IFABC = [1,0,0] can be represented as [I0, I1, I2] = [1/3, 1/3, 1/3].

We can confirm this using ( 30)


E. Franklin, B. Blackwell p29/50
ENGN4625/6625 2014 eBrick 1 - 30

I a  1 1 1  1 + 1 + 1  1


I abc =  I b  = 1 / 31 + 1 / 3h  + 1 / 3 h  = 1 / 31 + h 2 + h == 0 = 0
     2  
 I c  1  h  h 2  1 + h + h 2 == 0 0

This means we can replace the unbalanced fault current by three balanced source – the symmetric
components of IF. This then means that the impedance/admittance matrix is diagonal, and we only
need to analyse a single phase at a time, a great simplification. The circuit with a fault can be
redrawn as a circuit with three sets of balanced current sources. Although the circuit looks a lot
more complicated (9 new sources instead of one – remembering that the fault can be represented by
a current source) it is actually simpler to solve (though it probably will seem complicated still when
you first see the method!).

Figure 15: Single line fault, showing the actual circuit and the replacement of the fault by 3 sets of symmetric
component sources. Figure from Bergen 12.2

Solution to the problem comes from two important results of performing the SCT transformation:

1. we now have a balanced system (each phase is identical) and hence we can solve by per-
phase analysis
2. we can now use superposition and consider the three sequence networks separately

The three resulting networks (positive-sequence network, negative-sequence network, and zero-
sequence network) are shown in the figure below, noting that the system voltage source is a
positive sequence source. Zg is omitted from the positive and negative sequence networks because
E. Franklin, B. Blackwell p30/50
ENGN4625/6625 2014 eBrick 1 - 31

for a balanced Wye connected network there is no return current. By the same reasoning the zero
sequence network contains an open-circuit where the voltage source is because there cannot be any
zero sequence currents owing to it being a 3-wire Wye connection.

Figure 16: Per-phase sequence networks for the SLG problem (note ‘+’ used instead ‘1’, ‘-‘ instead of ‘2’) Figure from
Bergen 12.2

E. Franklin, B. Blackwell p31/50


ENGN4625/6625 2014 eBrick 1 - 32

Figure 17: The single phase analysis for the positive sequence, and (right) the equivalent circuit used to calculate the
voltage at the fault point (Va'g), based on the top line of the SCT from V012 to VABC. That equation states that the “A”
phase voltage is the sum of the three symmetric components – therefore they can be connected in series as above.

For the SLG fault considered here we need to realise how the three sequence networks need to be
combined. Once you become familiar with the individual sequence networks working out how to
combine them for a given fault condition is the main step which requires thinking about.

We can use superposition, or simply use the first line of the SCT to realise that

+&°à = +&°à
<
+ +&°à + +&°à .

h&á = h&á = h&á =


< âã
X
Also the SCT has yielded

âã
X
Hence the networks can in our case be combined by series connection with a common current

impedance: +&°à = há oá = 3oá


âã
flowing between the networks. Further, the current and voltage at the fault are related to the fault

So the networks are combined as shown in the figure below. The solution can be found via simple
circuit theory (you can do this to check it), and we find that:


há = 2
2ä o + o + o
3 à á

E. Franklin, B. Blackwell p32/50


ENGN4625/6625 2014 eBrick 1 - 33

6.4 Time domain analysis


So far we have considered phasor analysis. With the advent of real-time control, it is important that
a control algorithm feed back as quickly as possible, to keep the system stable. The evaluation of
phasor quantities needs at least one cycle of information, and preferably more to be accurate, which
can introduce delays in the loop and instability. Also even in post-event analysis, if things change
too quickly, the Fourier representation implied in phasor representation is not ideal.

This means that analysis and control system transforms working in real, instantaneous variables
(v(t), i(t) ) have advantages. The Clarke Transform is one such example.

6.4.1 The Clarke or α, β, 0 Transform for time domain analysis


Given that we don’t have phase information in real time, we cannot reproduce the nice property of

phase – reduces to just one component. This component would ideally be +t + +u ℎ + +v ℎ, but
the symmetric component transform that – for the most common situation – normal, balanced three

we can’t do the h operator without phase information. A close approximation is to use the 180°
component instead, and adjust the coefficients. Noting that VB + VC is at 180° to VA

+å = ® +t − ® +u + +v

To sense rotation direction, note that VB and VC interchange when rotation reverses, the difference
of these is at 90° to VA and will change sign when the rotation direction changes.

VC
-1/√2VB
Vβ VC

rotation -1/√6VB -1/√6VC


VA 1/√2VC VA
V
2/√6VA
VA,B,C
VB VB

Figure 18: Linear combinations of VABC (using only REAL coefficients) to obtain in-phase and quadrature (90°)
components

So make a similar linear combination based on the components at 90° (the difference +v − +u so
that normal three phase produces

• two equal components Vα, Vβ at 90° phase difference,


• Vα is in phase with VA.
• the phase of the second Vβ gives the rotation direction (90° lead is normal rotation)
• the total power in those components equates to the total three phase power (Power
Inv)
• one component V0 which is zero normally (same as zero sequence)

which when combined can represent all possible three phase phasor configurations (in the same
way as the symmetric component transform.

E. Franklin, B. Blackwell p33/50


ENGN4625/6625 2014 eBrick 1 - 34

œ√ƒ − √ƒ − √ƒ£ +å œ√ƒ − √ƒ − √ƒ£ +


› ¢ +æ Ú = ›› 0 − ¢ t
We define – = › 0 − √ √ ¢ √ √ ¢
˜+u ™
› ¢ +< › ¢ +v
such that Ù ( 38)

š√X √X
√X ¡ š√X √X √X ¡

is the Clarke transform from ABC to α, β, 0. Note that the order is different to V0,1,2 – the zero
sequence is last. This is one of the most common conventions in time domain transforms (and there
are very few conventions in transforms that are agreed by most authors)..

Clarke Transform 400

V 0 := 240 √2 V f := 50 Hz ω:= 2 π f Npts := 100 300

ind := 0… Npts t := 0.03∣ind / Npts sec 200


ϕ3 := 120 π/180 100
V_A
V A := V 0 cos( ωt ) V B := V 0 cos( ωt− ϕ3) 0 V_B

[V]
V C := V 0 cos(ωt − 2 ϕ3 ) -100 V_C

-200
plotxy ( t ,V A , t , V B , t , V C )
-300
Current Australian 3 ϕcolours
T -400
V ABC := [ V A , V B ,V C ] 0 0.01 0.01 0.02 0.02 0.03 0.03 0.04

[s]
2/ √6 − 1/ √6 − 1/ √6
ClarkeT :=
[ 0 − 1/ √2 1 / √2
1/ √3 1 / √3 1/ √3 ] 500

[ ]
1.000 0 0 400 V_A
T
ClarkeT ClarkeT = 0 1.000 0 Power Invariant V_alpha
300
0 0 1.000 V_beta
200 V_0
Note: In α ,β , 0 order (0 last, c.f. SCT where 0 is first).
100

VT := ClarkeT V ABC 0
[V]

-100
-200
T T T
plotxy ( t ,V A , t , VT 0, ind , t , VT 1, ind ,t ,VT 2, ind ) -300
-400
-500
0 0.01 0.01 0.02 0.02 0.03 0.03 0.04
[s]

Figure 19: Clarke transform operating on balanced three phase inputs VA,B,C (above) producing Vα (in phase) and Vβ (90°) and zero
sequence outputs. In the CompPad script above, VT is the transformed voltage vector, the plotting function plotxy requires column
vectors (hence the superscript T indicating transpose in the final line).

Unfortunately, there are many variations on the other details of these transformations, you have to
make sure you use a consistent set of matrices. We use the “modern” convention following the
EMTP theory book and “Power Systems Analysis” by Bergen and Vittal (although Bergen orders
0,α,β and 0,d,q). Apart from whether the order is α, β,0 (which we use) or (0,α,β – Bergen p227),
more modern texts (Bergen, EMTP theory) usually use the sign of the β row [0, -1/√2, 1/√2 ], that
we use, which means that for normal rotation the β component leads, and others (Fitzgerald, Bose,
MIT lecture notes) the opposite sign (and β lags). Also, there is a popular alternative that is not
power invariant, but preserves the voltage instead (contains numbers like 2/3). Differences between
authors are summarised in a footnote on the following page. Although the Clarke transform is
simple to write down, the true usefulness becomes clearer when we consider the DQ0 transform
below.

E. Franklin, B. Blackwell p34/50


ENGN4625/6625 2014 eBrick 1 - 35

6.5 The Park’s or D,Q,0 Transform for rotating machines


The DQ0 transform (idea by Blondel, development by
A-
Park et al.) is the equivalent of the Clark transform, but
transformed to a frame rotating with the rotor of a C+ B+
motor. This removes the oscillation in the α
component (becomes D - direct) and β (becomes Q N
quadrature) component. The Q or 90° component
grows as the rotor lags behind the rotating field, which Q Axis
increases the output power of the motor. This can be S
seen by reference to Figure 20, because the greatest
D Axis
force occurs when the magnet is between poles (lagging C-
B-
the field by 90°). If the magnet is at 0° relative to the A θ A+
phase electromagnet, which would be “6 O’clock” in A phase stator axis
the figure (not our usual convention), it is at a point Figure 20: Idea behind the
of minimum force, and if displaced, will return to three phase motor - a
rotating field is generated
the original position. Similarly, past 90°, the torque which pulls the magnet
starts to decrease. In this oversimplified diagram, around.
the 6 electromagnets are much narrower than in real motors, in which the electromagnets extend so
that the end of the A+ winding is close to the beginning of the A-.

The D (direct) axis is the axis of the magnet in this simple picture. In a large motor with an
electromagnet for the rotor, that electromagnet is complex, and has windings with equivalent axes
both in the D direction and the 90° or Q direction. Whether the Q direction is taken to lead or lag
the D direction again depends on conventions. We choose Q lagging the D direction in rotation
(although the component VQ leads VD), and that means with the rotor is lagging the rotating
magnetic field by 90° under load, the magnetic field lines up with the negative Q direction.

The DQO transform can be written as a simple 2x2 rotation on the α,β coordinates of the Clarke
transform combined with that transform.

+© cos − sin +å
•+ ’ = “ — • + ’
ç sin cos æ
( 39)

Alternatively it can be written as a time varying 3x3 as follows. Again, be wary of differences in
ordering of components and constant factors according to author2, and be careful to use consistent
ordering and the matching inverse. (You can retrieve the Clarke transform by putting θ = 0)

2
Bergen orders 0,d,q and is power invariant. EMTP orders d,q,0 and is power invariant. Fitzgerald orders d,q,0 and his
transformation is not power invariant (he lists one that is). Bose(p59), Kraus orders q,d,0, and is also not invariant.
Bergen takes θ as D axis relative to A axis(p222), and D leads Q, the q row is (p8-6)+ sin(θ -2π/3), +sin(θ+2π/3.
Krause takes θ as Q axis relative to A axis(p274), the Q axis leads, and the q row (p311) is + cos(θ -2π/3), +cos(θ+2π/3)
Fitzgerald θ as D axis relative to A axis(p271), the Q axis leads, and the q row (p271) is -sin(θ -2π/3), -sin(θ+2π/3)
EMTP theory θ as D axis relative to A phase axis(p8-21), Q axis lags, the q row is (p8-6)+ sin(θ -2π/3), +sin(θ+2π/3)
3
Actually θ is the angle of the rotor/Npp where Npp = the number of pole pairs (=1 in this case, one North, 1 South).
More on this in motors chapter.
E. Franklin, B. Blackwell p35/50
ENGN4625/6625 2014 eBrick 1 - 36

2] 2]
œ cos cos é − ê cos é + ê£
+© › 3 3 ¢
2› 2] 2] ¢ +t
+
Ù çÚ = è sin sin é − ê sin é + ê ¢ ˜+u ™
3› 3 3
¢ +v
DQ0
+< › 1 1 1
( 40)
› ¢
transform

š √2 √2 √2 ¡
where θ is the angle of rotation of the rotor3, and is typically ωt. This is suggestive of a constantly
changing transformation matrix, whose coefficients vary so as to put the amplitude of the in-phase
component into one output (VD), the quadrature component (not Vβ) into another (VQ) and the zero
sequence into the remaining one. Normal rotating three phase, with VA peaking at t=0 would
correspond to (VD=1, VQ=0, V0=0). The same waveform delay by 90° would produce (0,1,0), and a
further 90° → (-1,0,0). This is a nice alternative way to view the transform. Recall the picture in
Figure 18 of the Clarke transform as a linear combination (with real coefficients) of VABC. Ask

Is there a set of time-varying coefficients ô( ), ( ), _( )


the question –

so that ô( )+t + ( )+u + _( )+v removes the time variation of the Clarke transform Vαβ0?
Noting the constant power property of three phase +t + +u + +v = Const (typically 415V) the
coefficients are cosine functions like VA, VB.VC respectively - that is cos , cos ( − 2]⁄3),
cos ( + 2]⁄3) as in the top row of ( 40). Similarly for the second row (sin), and the last is just
the same as the zero sequence in the previous transformations. This is spelt out in more detail in
the example given in class, which is on the next page.
Although this results in a much more slowly varying output, as shown in Figure 21, in fact the
output is available with a very short delay (or latency) time – there is no need to average over a
cycle (or half cycle) to obtain the amplitude and phase. This is therefore highly suitable for high
performance motor control systems with fast response times, without risking instability – vector
control. The information required for control is available instant by instant, provided that the
variation of θ with time is known accurately. (Simulink induction motor control example :
http://www.mathworks.com.au/help/toolbox/physmod/powersys/ug/f4-10148.html)

Figure 21: Motor startup voltage waveforms, for a simple model of a motor. Current waveforms would be very similar (not shown).
The right figure shows the effect of the dq0 transform on current waveforms, making them much simpler. The D cpt (red) represents
the real power, Q (yellow) the reactive power. Both show an inertial lag during acceleration, which mainly changes the reactive
impedance during that time. The dashed lines show the current waveforms with the lag eliminated. Note that these are stator
currents. The most common application is for rotor currents (fictitious and actual).
This highlights one of the potential problems in application – how to obtain the angle (θ) of the rotor?

3
Actually θ is the angle of the rotor/Npp where Npp = the number of pole pairs (=1 in this case, one North, 1 South).
More on this in motors chapter.
E. Franklin, B. Blackwell p36/50
ENGN4625/6625 2014 eBrick 1 - 37

Rotary encoders are one mechanical solution. Various intelligent algorithms from phase locked loops to
neural networks can be used to estimate θ if encoder information is not available.

E. Franklin, B. Blackwell p37/50


ENGN4625/6625 2014 eBrick 1 - 38

Example:

We can obtain this transform in a different way if we ask the following questions (an expanded
form of the question asked on the previous page):

How can we obtain the instantaneous voltage from a three phase system?

Answer – compute +öX = µ+t + +u + +v

This reflects the amplitude of all three components, but loses phase information (Error! Reference
source not found.). This may be useful but sometimes we must have phase information.

How can we obtain both magnitude and phase information quickly from a 3 phase system?

Answer – replace VA2 above with VA cos(ωt); that is one of the terms is the actual VA, including
any phase shifts, and the other is a copy of VA that is fixed in phase, and so on for VB2 and VC2.

Figure 22: Three phase system showing a phase jump at 20ms, a decrease in magnitude from 40-60ms and a frequency
shift (equivalent to a linear increase in phase) from 100-120ms. Magnitude is quickly and accurately measured by
√(VA2+VB2+VC2), but phase variations are ignored. The right graph shows that replacing VA2 by VA cos(ωt) etc., a
quadrature pair of signals VD, VQ (dashed green and yellow respectively) are generated, returning instantaneous values of
both amplitude (√(VD2 + VQ2)) and phase (atan(VD/VQ)).

This is just the first line of the DQ0 transform, replacing θ by ωt:

VD = CX [ +t cos + +u cos D − F + +v cos D + F],


‚ ‚
X X
VQ = ….. etc.

Now the in-phase component is in VD, and the 90 phase component is in VQ. This can be related to
the answer to the first question (VM3): VD = VM3 cos(ϕ) and VQ = VM3 sin(ϕ), where ϕ is the phase
delay of the voltage set being measured. Figure 23 shows a typical real time controller using these
techniques. The coordinate transforms can be performed by high speed digital signal processors
(DSP) such as the TMS320C30 .

E. Franklin, B. Blackwell p38/50


ENGN4625/6625 2014 eBrick 1 - 39

Figure 23: Block diagram of "Vector Controller" using the dq0 transform for high speed estimation of motor variables,
and fast responding control (Wikimedia commons)

Another simplification provided by this transformation is that the inductances, which depend on the
shaft rotation, are made independent of time by this transformation. Considering Figure 20, the
iron in the rotor would strongly couple the A+ coil to the A- coil, in the orientation shown,
increasing the self inductance of the A coil set. 90° later, this coupling would disappear (and
would cross-couple B and C). We therefore need to use V = d/dt(LI) = I dL/dt + L dI/dt in KVL.
The transformation to the rotating rotor frame removes this dependence to a reasonably good
approximation. How this happens is beyond the scope of this course.

The Kron transform refers variables to the frame of the rotating magnetic field generated by the
stator windings, instead of the rotor frame. In equilibrium, for a synchronous motor, the two frames
rotate together. Differences between the two frames occurs mainly during start-up and load
changes, so the Kron approach is used for analysis of such dynamic problems.

There are many more concepts and terms associated with motor/generator theory that we will cover
in a little more detail in the chapter on motors.

E. Franklin, B. Blackwell p39/50


ENGN4625/6625 2014 eBrick 1 - 40

7 Per Unit Notation


Details of transformers and motors are often described more simply using impedances, voltages,
currents and power in “per-unit” notation, where the quantity is expressed as a fraction of a base
unit, usually the nominal or “identification plate” power of the device. For example if a transformer
designed for 110kVA power transmission at a nominal voltage of 11kV, the primary current would
be 100A, and the ratio V/I is 100Ω. We might choose this as the “Base Impedance” to which we
refer other impedances, such as the series (or leakage) inductance of the windings. Say the leakage
inductance had a reactance of 5 j Ω, then we would call that 0.05 per unit, or 5%.

The main advantage of using per-unit notation is when transformers are involved (or fault currents
are being calculated in power distribution systems). In this case impedance and voltage drop
calculations can be done without considering the transformer ratio and a complex power system is
somewhat simplified and easier to inspect with meaningful interpretation. In addition, often the %
voltage drop is more meaningful than the actual. When the real units are required, the actual units,
taking into account the turns ratio, are multiplied in.

These are not recommended for beginners, because a lot of information is lost in the process, so it is
harder to find mistakes. However any electrical power engineer does need to know the notation.
Per unit notation is the standard in power systems analysis.

Since the four basic system variables - power, voltage, current and impedance (or admittance) – are
linked via two equations S = VI* and V = IZ we can only arbitrarily choose two base quantities and
the other base units will follow. It is normal for power systems to choose power and voltage as the
chosen base units and then calculate the base current and base impedance quantities, but this can be
different depending upon the situation. Remember that base quantities are magnitudes only, so the
variables all still retain the same relative complex values in per units notation.

i +
So if for example we choose base power SB and a base voltage VB (which might be the rated power
of our load and the system voltage) then we can use hu = uä+ and ou = uäh . We then simply
u u

convert all of our real variable (which are phasors) to per unit values via +÷ø = +ä+ etc. and then
u
all analysis can take place as per normal.

7.1.1 Example using per-unit notation:


The impedance of a 350 kVA 11kV/415V transformer is 4.5%, the X:R ratio is 4.0 and supplies
rated load at 0.707 power factor lagging. Calculate the secondary voltage for rated load on the
assumption that there is no voltage regulation in the source. (These transformer concepts
(impedance, current and voltage ratios, and regulation) are treated later in the transformer and
motors section of the course)

Translation! 350kVA means the total kVA over three windings (as it doesn’t say otherwise).
Impedance in this type of problem refers to the Thevenin equivalent series impedance, sometimes
called short circuit impedance. It is usually mainly leakage inductance, deliberately included
during manufacture to limit short circuit (fault) currents. It is not stated whether the transformer is
delta or wye, but we can assume that the primary and secondary have the same topology. (One of
the virtues of per unit notation is that you can get the answer without knowing all the details, such
E. Franklin, B. Blackwell p40/50
ENGN4625/6625 2014 eBrick 1 - 41

as is it delta or wye?). No voltage regulation in the source means that the source voltage doesn’t
change when connected to this circuit. The X:R ratio is given because the phasor relationship
between the terminal voltage and the voltage drop across the impedance matter, as they
add/subtract vectorially. Note that the problem already lets us know that we need to use per unit
notation (with impedance being measured as a %).

If we call the rated transformer voltage v := 1 per unit and the current
− acos (0.7071) j
i L := 1 e = 0.7071− 0.7071 j , then the impedance is
j atan(4)
z L := 0.045 e = 0.01091+ 0.04366 j
Set the primary voltage v inp := 1 , the the secondary voltage is
v LD := v inp − i L z L = 0.9614− 0.02315 j
So voltage drop is v drop := ∣v inp∣− ∣v LD∣= 0.03831 per unit.

The secondary voltage magnitude in real units is thus 0.962*415 = 399V

It is worth noting that a change of base in per unit system analysis is simple. It is typically required
when analysing at systems with several components. A conversion to a new per unit base can be
done simply without having to go back to real quantities first. You can probably work out
yourselves how to do this… we won’t cover it in this course.

E. Franklin, B. Blackwell p41/50


ENGN4625/6625 2014 eBrick 1 - 42

8 Fourier Series & Harmonics


Any periodic waveform with period T (= 1/f1 ) can be represented exactly by an infinite sum of sine
and cosine functions having frequencies which are integer multiples of the original waveform
frequency f1. The series is known as a fourier series.

In power electronics, waveforms are mostly periodic and are symmetric is various respects. This
allows the use of Fourier series in place of integrals (if f(t) is periodic), and allows simplification by
the use of the trigonometric form (i.e. cosine, sin instead of ejθ), by exploiting symmetries. In power
electronics it is extremely useful to consider a waveform in terms of its fourier series.

We will see that the Fourier series for the input current to a 3 phase rectifier needs only cosine
terms, and of those, only odd harmonics (f = nf1, where f1 is the fundamental (50Hz) and n is odd).

8.1 Fourier series definition


A periodic waveform f(t) can be represented by the fourier series as follows:

T = + ùúý ú cos -_ + ûú sin -_


ü
< ( 41)

where < = 0 þ ¦ô /T T , and

= ‚ E< T cos -_ . _

ú ( 42)

ûú = ‚ E< T sin -_ . _

( 43)

8.1.1 Symmetry
The symmetry of the periodic function (symmetry about the y-axis) can be taken advantage of to
simplify the fourier series coefficients further since only some are non-zero. The following table
(taken from Mohan Chapter 3) shows the coefficients for common symmetries and their definitions:

E. Franklin, B. Blackwell p42/50


ENGN4625/6625 2014 eBrick 1 - 43

We can see for example that with even symmetry,

• all sin terms will go to zero, as the product will be a symmetric odd function
• for the remainder, we only need to integrate to the right of the origin.

Half wave symmetry means that the wave repeats with inversion every half cycle. (like the current
in a full wave rectifier (confusing ‼).

And another commonly useful symmetry is even quarter wave symmetry (which is just even and
half wave combined). – waveforms with this symmetry only have one in four of the components
non-zero.

8.1.2 Example - fourier series coefficients for a square wave


Consider a square wave with zero average and amplitude 1. If the waveform is drawn with odd
symmetry then the an coefficients are all 0.

2π wt

2 ‚ 2 −cos -_

ûú = 8 sin -_ . _ =
] < ] - <

2
= − cos -] + 1
-]
E. Franklin, B. Blackwell p43/50
4
ENGN4625/6625 2014 eBrick 1 - 44

ûú = T/0 - = 1,3,5,7 … ûú = 0 T/0 - = 2,4,6,8 …


-]
Hence the square wave function can be written as:

4 4 4 4
T = sin _ + sin 3_ + sin 5_ + sin 7_ +⋯
π 3π 5π 7π
These first four terms are plotted below, along with the sum of these terms. The square wave
waveform can clearly be seen taking shape. These harmonic components of the

1.5

0.5

0
0 5 10 15 20 25 30 35
-0.5

-1

-1.5

8.2 Consideration of fourier analysis for power conversion circuits


We will see later in the course that many switch mode power conversion technologies generate
voltage and current waveforms which, if they are not filtered, are not the ideal or desired waveform.
Likewise currents drawn from mains power supply often contain many undesired harmonic
components. Fourier analysis helps enormously to understand these situations and allows the design
of filters for wave smoothing or circuits for removal of harmonics. As an example of this consider a
DC-AC power conversion switching circuit intended to form a sinewave output. If the primitive
output is a square wave then we can determine the sinusoidal components that it is composed of, ,
as analysed above. If we were to apply a 2nd order filtering function to the fourier series above such
that the filter gain is proportional to 1/w2 then the sum of the filtered fourier series components is
dominated by the fundamental harmonic and the waveform becomes very close to a pure sinewave.
The filtered waveforms (first four terms and the sum of these terms) is shown here.

1.5

0.5

0
0 5 10 15 20 25 30 35
-0.5

-1

-1.5

E. Franklin, B. Blackwell p44/50


ENGN4625/6625 2014 eBrick 1 - 45

8.3 Fourier series of some common waveforms

sin -_
äY
T =
«t
‚Y úY
úý ,X,Z…

ü
T = sin -_
Ît
‚ úý ,X,Z… ú

ü
T = + sin _ − cos -_
t t t
‚ ‚ úý ,Î,ƒ… ú
Y

ü
T = − cos -_
t Ît
‚ ‚ úý ,Î,ƒ… ú
Y

8.4 Power factor revisited (Non-linear Power Factor) & harmonic distortion
We previously introduced power factor, defined as the cosine of the phase angle between voltage
and current, or equivalently the angle of the complex power variable. This is otherwise known as
the Displacement Power Factor (DPF) – displacement referring to the displacement in time or phase
difference.

However, in power electronics we often have a sine wave voltage supply and a non-linear circuit
which results in the current waveform consisting of the fundamental (with the same frequency as
the voltage waveform) plus various harmonics. Recalling that each of the harmonics are orthogonal
to the fundamental it follows then that only the fundamental current waveform draws any real
average power from the source; current is drawn from the supply for all other harmonics but with
no net power transfer (analogous with current being drawn to supply reactive power only). The true
Power Factor or Non-Linear Power Factor (NLPF) (sometimes even just referred to as Power
Factor) is then the ratio of the fundamental to the total current multiplied by the Displacement
Power Factor.

s4j = cos

h¬ör
U 4j = . s4j
h¬ör

Another commonly used term is the Total Harmonic Distortion (THD). It is a measure of the
combined magnitude of the harmonics as a fraction of the fundamental waveform… so ideally it
E. Franklin, B. Blackwell p45/50
ENGN4625/6625 2014 eBrick 1 - 46

would be close to 0. THD is defined by the squares of voltage or current waveforms (since it can
be related to power) as follows (values are RMS values):

h + hX + hÎ + ⋯ + hü
cÓs = è
h

Since (from RSS theory) we know that h = h + h + hX + hÎ + ⋯ + hü , we can write:

h + h + hX + hÎ + ⋯ + hü h
cÓs + 1 = =
h h

And thus easily re-write the expression for NLPF in terms of THD:

1
U 4j = . s4j
√1 + cÓs

9 Computer Simulation – LTC Spice


Circuit simulators such as the University of California’s (UCB) SPICE4 have been used in small-
signal analysis of linear circuits. Mesh equations in complex variables are automatically
assembled, and solved using linear matrix solutions. However Power electronics requires a non-
linear simulator.

There are two simulators relevant to this course –

SPICE – which has


detailed device models and
sophisticated numerical integration
control, but limited transmission
lines, and no three phase plant.

ATP – which has simple


switches, but detailed models of
power systems plant and
transmission equipment, and
handles transients in detail.

ATP (the Alternative Transients


Program - http://www.emtp.org/) is accompanied by a theory book (The EMTP Theory book – on
web ) which is a compendium of a wide range of power systems theory relevant to analysis and
simulation, particularly ATP.

ATP will only be available this year to Master’s students who are willing to tackle this powerful but
formidable tool.

4 Figure 24: Screen shot of the ATP simulation of various generators and
Simulation Program with Integrated Circuit Emphasis
motors connected by three phase buses.
E. Franklin, B. Blackwell p46/50
ENGN4625/6625 2014 eBrick 1 - 47

Spice can be used to simulate power electronics (although originally for ICs) LTSpice is more
suitable than pSpice, as there are no limitations, and optimised for Power. Power circuits are
usually non-linear, so transient solutions (time-dependent solutions for V and I) are frequently
required.

E. Franklin, B. Blackwell p47/50


ENGN4625/6625 2014 eBrick 1 - 48

9.1.1 Hints for running LTC Spice


• Need to complete the circuit and include a ground (G)
• Voltage sources/batteries are all in one – use the advanced tab to choose sin or pulse etc.
• To see time dependent waveforms, or to model non-linear devices (e.g. diodes), need
Simulate/Edit Simulation Command /– Transient mode tab. (stop time is all you need)
• Probes are automatic – change to current probe if inside a component.
• Power probe (thermometer) – use Alt-key and hover on component – generates a new plot.
• Average power – Ctl- click on plot title to calculate average and RMS, Clt-click on power
plot to get average power and energy.
• Adjust plot range to select average interval, which should be a whole number of cycles.
• FFT under Plot/View Menu – make sure you include between 8 and 100 cycles of data.

9.2 pSpice (pSp) and LTSpice (LT) hints and comparisons.


• We mainly use “Transient Analysis” (Under “Analysis” find Setup/Transient) because power
electronics circuits are usually non-linear (rectifiers), the signals are not “small”, or the start-up
transients are significant (capacitor charge-up transient). The usual “AC” or “small-signal”
analysis although very fast, is not valid in these cases.
• This means we should use sources (VSIN, VPULSE) which are designed for transient analysis,
not VAC. In LTSpice, these are the same icon, the transient version is accessed under advanced.
• Drop the Voltage probes on nodes to see the voltage. The differential probe (Under Markers –
Mark voltage differential lets you plot the voltage between points A and B – two probes + and -
). Current probes need to be dropped in an unambiguous place (i.e. not right at a node).

E. Franklin, B. Blackwell p48/50


ENGN4625/6625 2014 eBrick 1 - 49

9.2.1 Accuracy:
1) Most commonly, the values are accurate, but the point spacing on the graph may undersample
the data, giving the wrong impression – e.g. jagged waveform. Use Maximum Timestep (LT)
or Step Ceiling (pSp) under Setup/Transient to request finer steps (the temporary output file will
grow…)
2) .OPTIONS: Typically Power Electronics problems require relaxed accuracy – calculating a
100Amp current to 1picoAmp accuracy is neither necessary nor efficient. Two typical tuning
adjustments are abstol=1e-6 and gmin=1e-6. This is under Setup/Options(pSp) and
Simulate/Control Panel/Spice(LT). In LT, most of these setting are not saved, so it is better to
use Spice directive text (next to comments), they will stay with the circuit, and be easily visible
so you don’t forget.
Example: .options abstol=1e-6 gmin=1e-6 ; LTSpice – relaxed accuracy.

• Be careful relaxing accuracy if you are using low power devices as well – such as a micropower
comparator or Opamp. They will probably operate on much smaller currents. The alternative is
to use behavioural voltage/current sources (bv). As a last resort, you can try reltol=0.003 or
0.01, but check that the results don’t change appreciably with reltol (except you would expect
the computation to run faster – that is the point!
• No spaces allowed in values: (e.g. 10 Ohm gives an error “Missing value”: instead use 10.Ohm -
decimal point separates the “Oh” from the zero clearly (or 10Ohm or 10R or 10ohm) ) Also
need to spell MEG 10Mohm means 10 milliohm, similarly 7MegaHz for frequency (MegaHz
is OK at least in LTSpice).
• You need an earth somewhere e.g GND_EARTH.
• AC Sweep analysis does not need a particular input voltage source, except that it cannot be
0Volts.

9.2.2 Saving setup time in “Probe”:


1. Use voltage and Current markers in the schematic editor

2. If you set up complicated analysis, and wish to repeat that for a number of circuit changes, you
can check “Restore Last Probe session” under Analysis/Probe Setup in the Schematic editor.
Downside can be error messages about “missing” traces when changing files.

3. For LTSpice, you can save a plot file (.plt) with the plot setup you like (variables, scales etc). If
it has the same first name as the .asc main file, it will automatically determine what is plotted,
scales etc.

9.2.3 Libraries:
• The pSpice library browser is tricky to use. This works: put * in the left box (Part name). Put
*fet* in the right (Description search) and press the search button, then click on the returned
(FET) part names to see the description. There are only a few transistors, JFETs, two
MOSFETs one IGBT and one SCR (2N1595). The 1N4002 diode model (D1N4002) is quite
good.

• The LTSPice library has fewer components at first sight, but many are multi-function – e.g.
there are >30 NPN transistors in the properties of the NPN device. (and many of their

E. Franklin, B. Blackwell p49/50


ENGN4625/6625 2014 eBrick 1 - 50

proprietary devices) . You can also easily use manufacturer’s library files. Some .sub files can
be fond in the Simulations handout folder to show the syntax, and they should be put in the
directory c:/Program Files/LTC/SwCADIII/lib/sub. (see the example PWM_Half_Single.asc)
.sub files are the simplest, they only have electrical data, and need to be attached to an image.
The generic devices in lib/sym/misc (e.g. NIGBT) are easy to attach models to, just put the
model name (not filename) in the SpiceModel row in the properties (right click symbol). For
the components that have multiple choices (e.g. Diode), you need to right click on the diode
model name (not the symbol) to be able to type you own model name in. pSpice should also be
able to accept this type of file, but I haven’t worked out how yet, and you will quickly run into
the component count limit.

• Generic devices: pSpice has “BREAK” devices (very simple model) and LTSPice has a
skeleton IGBT symbols to which you can attach properties or .sub circuits (see the example
PWM_Half_Single.asc)

RMS: The simplest is the RMS() function in add/Trace. However if there is time dependence
(transients), then to calculate the final or “steady state” value, you must either remove data from
early times and ensure that enough time has elapsed, or use a one cycle average, which
converges much more rapidly. For example sqrt(avgx(I(R1)*I(R1),20ms)) is a 50Hz 1 cycle
(i.e. over 20ms: avgx is a running average) - this gives the RMS current in R1. To discount data
early times, use the “No Print Delay” field in pSpice (Setup/Transient) or the “Time to start
saving data” under LTSpice. Note: in LTSpice, RMS and Average under ctl-Left mouse (trace
title)

• Similarly, average power can be calculated as sqrt(avgx(I(R1)*(V(R1:2)-V(R1:1)), 20ms)). You


can define a macro under pSpice (probe/Trace/Macros) for these (don’t forget to save, or just
keep the macro in a file and cut and paste when needed) example:

avgpwr(a) = avgx((v(a:1)-v(a:2))*i(a),20ms).
rms20(a) = sqrt(avgx(a*a,20ms))
Note: the default file type is “probe utility” file .prb, which I tend to delete often……careful.

• Power in LTSpice – Alt left click brings up a “thermometer” icon, and adds the VI product to the
plotted quantities. Doesn’t work in AC sweep mode. Also, there doesn’t seem to be a running
average, (avgx in pSpice), but you can “low pass filter” power by making a b voltage source
with I and V in the expression, and feeding a filter.

• Bode (dB/phase vs freq) or Nyquist (Complex plane, freq is parameter) available in LTSpice –left
click on axis.

• XY plots in LTCSpice left click on x axis to change what is plotted there.

E. Franklin, B. Blackwell p50/50

Potrebbero piacerti anche