Download as pdf or txt
Download as pdf or txt
You are on page 1of 162

UNIVERSITY OF THE WITWATERSRAND

SCHOOL OF ELECTRICAL
AND
INFORMATION ENGINEERING

POWER SYSTEMS

ELEN5008A

©Professor John Van Coller and Dr Salman Minhas, 2024


Notes:

2 ELEN5008A – Power Systems


CHAPTER 1: BASIC CONCEPTS OF TRANSMISSION LINES

1.1 Introduction

Transmission lines transfer power at high voltages (above 132 kV in South Africa) from
generation (power stations) to loads (cities).

In this chapter we consider only AC transmission lines and hence all analysis involves phasors
where inductance and capacitance play a major role.

One of the most important equations for transmission lines relates the active power (MW) flow
along a transmission line P to the angle between the voltage phasors at the ends of the
transmission line δ.

For overhead transmission lines (modelled as a series inductive reactance only - series resistance
is much smaller so it is assumed to be zero)

V V
P = S R sin
Xs

where VS is the rms magnitude of the sending-end line voltage phasor, in V


VR is the rms magnitude of the receiving-end line voltage phasor, in V
X s is the series inductive reactance, in Ω
 is the angle between the sending-end voltage phasor and the receiving-end
voltage phasor, in degrees

The above equation is derived from the equation for the complex power flow per phase from bus
S (sending end) to bus R (receiving end)

SSR = PSR + jQSR = VS I S

- where VS is now the sending-end rms phase-neutral voltage phasor and I S is the complex
conjugate of the sending-end phase current (the complex conjugate is required because we want
to subtract the current phasor phase angle from the voltage phasor phase angle whereas if we did
not specify the complex conjugate of the current phasor, the phase angles of the voltage and
current phasors would be added, following the normal rules of complex number arithmetic).

V − VR
But IS = S
Z

ELEN5008A - Power Systems 3


Substituting into the previous equation
 V −V  V 2 − V V (  −  )
SSR = PSR + jQSR = VS I S = VS  S = S
* R S R S R
 Z  Rs − jX s
 
2
VS − VS VR cos(  S −  R ) − j VS VR sin(  S −  R )
=
Rs − jX s
 V 2 − V V cos(  −  ) − j V V sin(  −  )  R + jX
 S
 S R S R S R S R  s
 s ( )
=
2 2
Rs + X s

We define the power angle,  ( =  S −  R ) .

We calculate the active and reactive power flow per phase by equating the real and imaginary
components of the above equation

( VS )
Rs Xs
PSR = VS − VR cos  + VS VR sin 
Rs 2 + X s2 Rs 2 + X s2

( VS )
Rs Xs
QSR = − VS VR sin  + VS − VR cos 
Rs 2 + X s2 Rs 2 + X s2

The three-phase power flows are three times the above. Alternatively, the rms phase-neutral
voltages can be replaced with the rms line voltages and then we do not need to multiply by three.

Later on we work with admittances, so to convert series impedance Z s (assuming an inductive


series element) to series admittance Ys
Z s = Rs + jX s
Ys = Gs + jBs

-where Gs is the series conductance, in Mho


Bs is the series susceptance, in Mho

1 1 R − jX s Rs Xs
Thus Ys = = = s = −j = Gs + jBs
Z s Rs + jX s R 2 + X 2 R 2 + X 2 R 2
+ X 2
s s s s s s

Rs
Thus Gs =
Rs 2 + X s 2
Xs
Bs = −
Rs 2 + X s 2

4 ELEN5008A – Power Systems


Thus the previous active power flow equation becomes

PSR = Gs VS ( VS − VR cos  ) − Bs VS VR sin 


= Gs VS ( VS − VR cos ( S − R )) − Bs VS VR sin ( S − R )

Similarly the previous reactive power flow equation becomes

QSR = − Gs VS VR sin  − Bs VS ( VS − VR cos  )


= − Gs VS VR sin ( S −  R ) − Bs VS ( VS − VR cos( S −  R ))

For overhead transmission lines we assume (but definitely not so for cables)

X s  Rs
1
- which implies that we can ignore Rs (set it to zero) and hence Gs = 0 and Bs = − ( Bs is
X s
negative for an inductive series element).

Thus an alternative form of the active power flow equation for when bus k is connected to N
adjacent buses is ( Bkn is negative for an inductive series element)
N
Pk =  − Bkn Vk Vn sin(  k −  n )
n =1,n  k

Also an alternative form of the reactive power flow equation for when bus k is connected to N
adjacent buses is ( Bkn is negative for an inductive series element)
N
Qk =  − Bkn Vk ( Vk − Vn cos(  k −  n ))
n =1,n  k

V V V V
The above equations become PSR = S R sin  = S R sin(  S −  R )
Xs Xs

VS ( VS − VR cos  ) VS ( VS − VR cos(  S −  R ))
QSR = =
Xs Xs

Theoretically the maximum steady-state active power that can be transferred is (  =  / 2 or


90 )
V V
Pmax = S R
Xs

In practice, the steady-state value of  is not allowed to exceed approximately 35 to provide
an adequate safety margin for disturbances (see later).

ELEN5008A - Power Systems 5


Also, to increase the active power transfer limit along a particular transmission line corridor the
options are
- increase the line voltage (implies larger towers and longer insulator strings)
- reduce the line series inductive reactance (implies a series capacitor or reducing the
distances between phase conductors)

The active power transfer limit may also be reduced because of


- conductor overtemperatures
- unacceptably low receiving-end voltages

Another important impedance associated with the transmission line is the characteristic (surge)
impedance, Z o , where
XL' L'
Zo = =
YC ' C'

- where X L ' is the series inductive reactance per unit length, in Ω/m
YC ' is the shunt susceptance per unit length, in S/m
L' is the series inductance per unit length, in H/m
C' is the shunt capacitance per unit length, in F/m

Shunt capacitors can provide voltage support if the receiving-end voltage drops too low and
shunt inductors (line reactors) can reduce the voltage if the receiving-end voltage is too high.

The optimum active power flow that avoids the need for shunt capacitors and shunt reactors is
known as the Surge Impedance Loading (SIL)
2
VL
SIL =
Zo

- where VL is the rms magnitude of the line voltage, in V


Z o is the characteristic (surge) impedance, in Ω

The SIL is in MW if the rms line voltage is expressed in kV. For example, for a 400 kV line with
a characteristic impedance of 300 Ω
2
400 kV
SIL = = 533 MW
300 

A line loaded at its SIL has no reactive power flowing either into or out of the line both at the
sending end and at the receiving end (the line provides its own reactive power requirements).

A line loaded at its SIL is considered to be loaded at its ‘natural loading’.

In practice, the active power flow usually differs from the natural loading, and then

If PL  SIL → VR  VS
If PL  SIL → VR  VS

6 ELEN5008A – Power Systems


CHAPTER 2: TRANSMISSION LINE PARAMETERS

2.1 Inductance per unit length

2.1.1 Internal and external flux linkage per unit length

Consider a solid cylindrical conductor of radius r and assume that the current density is constant
within the conductor and the conductor has absolute permeability  ( =  r o )

If the total current flowing in the conductor is I , the current flowing within a radius x is
2
x
Ix = I
r2

Ix x
The flux density at radius x ( x  r ) is Bx =  H x =  = I
2 x 2 r 2

The flux per unit length in the cylindrical element of thickness dx is therefore

x
d ' = Bx dx =  I dx
2 r 2
2
x
This links with only of the total current in the conductor and therefore the flux linkage
r2
contribution per unit length of this cylindrical element of thickness dx is reduced to

2 3
x x
d' = Bx dx =  I dx
r2 2 r 4

The total internal flux linkage per unit length up to radius r is (uniform current density)

r r
I I I
int ' =  d  ' = x
3
dx =
4
( r −0 ) =
2 r 4 8 r 4 8
0 0

Note that when skin effect is taken into account, the above value is multiplied by k s (see later).

ELEN5008A - Power Systems 7


We now consider the external flux linkage per unit length up to point P

P (arbitrary point in space)


dx D
x

I
The flux density at radius x ( x  r ) is Bx = o H x = o
2 x

The flux per unit length in the cylindrical element of thickness dx is

d ' = Bx dx

This links with all the current in the conductor and therefore the flux linkage per unit length
contribution of this cylindrical element of thickness dx is
I
d ext ' = Bx dx = o dx
2 x

The total external flux linkage per unit length up to point P is therefore
D o I D 1 o I  D 
ext ' =  d ext ' =
2  x
dx = ln  
2  r 
r r

2.1.2 Total flux linkage per unit length and self-inductance per unit length with flux linkage
measured up to point P

Adding both the internal and the external flux linkages per unit length up to point P
(  =  r o )

r  
     I
    D   o I   I
4 ) + ln  D   = o ln 
 ' = I  o r + o ln   = ln( e  
D  = o ln  D 
 8 2  r   2   r   2  
− r  2  r' 
   4 
 r( e )

' D −7  D 
Thus L' = = o ln   = 2  10 ln   in H/m
I 2  r'   r' 

− r
- where r' = r( e 4 ) is the Geometric Mean Radius (GMR) of the single conductor (radius of
conductor with no internal flux that would have the same total flux linkage for the same current).

8 ELEN5008A – Power Systems


Thus if we replace r with r' in our calculation we need consider only the external flux linkage.

A conductor with a high relative permeability (steel) has a very small GMR - to replace the high
flux density inside the steel conductor, it is necessary to replace the actual conductor with a
conductor with a very small radius (trying to achieve in air (high flux densities outside the very
small conductor) with what occurs inside the steel (high flux densities due to high permeability).

2.1.3 Flux linkage per unit length in multi-conductor systems

Consider a group of n conductors, four shown below, where the sum of the currents in the
conductors is zero (no flux encircling all conductors - makes the total flux linkage finite)

P (arbitrary point in space)

D3P
D4P
3 D2P
D34 D
D23 2 1P

4 D12
D41

The flux linkage per unit length up to point P associated with conductor 1 , caused by the
current I 1 flowing in conductor 1 is
D 
1P1 ' = 2  10 −7 I1ln  1P 
 
 r'1 

Similarly the flux linkage per unit length up to point P associated with conductor 1 , caused by
the current I 2 flowing in conductor 2 is (flux encircling conductor 2 that also encircles
conductor 1 )(flux encircling conductor 2 within distance D12 does not link with conductor 1 )

D 
1P2 ' = 2  10 −7 I 2ln  2P 
 
 D12 

Thus the total flux linkage per unit length up to point P associated with conductor 1 , caused by
all n currents, is

 D  D  D  D 
1P ' = 2  10 −7  I1ln  1P  + I 2ln  2P  + I 3ln  3P  + ... + I nln  nP 
  r'   D   D   D 
  1   12   13   1n 

ELEN5008A - Power Systems 9


−7
  1   1   1   1 
= 2  10  I1ln   + I 2ln   + I 3ln   + ... + I nln  
  r'  D  D   D 
  1  12   13   1n  
+ 2  10−7 ( I1ln( D1P ) + I 2ln( D2P ) + I3ln( D3P ) + ... + I nln( DnP ))

Substituting just in the second term I n = − ( I1 + I 2 + ... + I n −1 )

  1   1   1   1 
1P ' = 2  10 −7  I1ln   + I 2ln   + I 3ln   + ... + I nln  
  r'  D  D   D 
  1  12   13   1n  
+2  10 −7 ( I1ln( D1P ) + I 2ln( D2P ) + I 3ln( D3P ) + ... + I n −1ln( D( n −1 )P ))

−2  10−7 ( I1ln( DnP ) + I 2ln( DnP ) + I 3ln( DnP ) + ... + I n−1ln( DnP ))


 1   1   1   1 
Or 1P ' = 2  10 −7  I1ln  + I 2ln   + I 3ln   + ... + I nln  
  r'  D  D   D 
  1  12   13   1n  
−7
 D  D  D   D( n −1 )P  
+2  10  I1ln  1P  + I 2ln  2P  + I 3ln  3P  + ... + I n −1ln  
 D  D  D   D 
  nP   nP   nP   nP  

If arbitrary point P is moved to an infinite distance away

D1P D D D( n −1 )P
→ 2P → 3P → ... → →1
DnP DnP DnP DnP

- and since ln( 1 ) = 0 , the whole of the second term in the previous equation becomes zero.

Thus the total flux linkage per unit length associated with conductor 1 is then (requires the sum
of currents in all the conductors to be zero)

  1   1   1   1 
1 ' = 2  10 −7  I1ln   + I 2ln   + I 3ln   + ... + I nln  
  r'  D  D   D 
  1  12   13   1n  

10 ELEN5008A – Power Systems


2.1.4 Conductor bundles

Consider a line made up of two conductors where each conductor is a conductor bundle
- a conductor bundle replaces a single conductor with a group of subconductors that are
electrically in parallel
- distances between subconductors are denoted by Dij
- assume conductor bundle X carries current I in one direction and conductor bundle Y
carries the same current I in the opposite direction (so that the sum of the currents in all the
conductors will be zero)
- assume conductor bundle X consists of n identical subconductors each carrying a current
I / n and conductor bundle Y consists of m identical subconductors each carrying a
current I / m
- to make the notation more consistent, the Geometric Mean Radius (GMR) of subconductor
i , ri ' will be denoted by Dii

The total flux linkage per unit length associated with subconductor 1 in conductor bundle X
due to the currents in the subconductors in both conductor bundle X and conductor bundle Y is
then (subscript numbers with dashes are for the subconductors in conductor bundle Y )(requires
sum of currents in all the subconductors to be zero)

I  1   1   1  I   1   1   1 
1 ' = 2  10 −7   ln   + ln   + ... + ln    −  ln   + ln   + ... + ln  
 n   D11  D 
 
 D  m   D
 D12'
D 
  12  1n    11'    1m' 

mD 
11'  D12'  ...  D1m' 
Or 1 ' = 2  10 −7 I ln  in Webers/m
 n D  D  ...  D 
 11 12 1n 

The self-inductance per unit length of subconductor 1 in conductor bundle X is therefore

1 '  m D  D  ...  D 
L1 ' = = 2  10 −7 n ln  11' 12' 1m' 
I/n  n D11  D12  ...  D1n 
 

m 
−7  ( D11'  D12'  ...  D1m' )n 
= 2  10 ln   in H/m
 D11  D12  ...  D1n 
 

The average self-inductance per unit length of a subconductor in conductor bundle X is

L ' + L2 ' + ... + Ln '


Lav ' = 1
n

ELEN5008A - Power Systems 11


Since there are n subconductors in parallel in conductor bundle X , the self-inductance per unit
length of conductor bundle X is approximately

L ' L ' + L2 ' + ... + Ln '


LX ' = av = 1
n n2
 mn ( D  D  ...  D )( D  D  ...  D )
2m' )...( Dn1'  Dn2'  ...  Dnm'
= 2  10 −7
ln  11' 12' 1m' 21' 22' 
 n2 ( D  D  ...  D )( D  D  ...  D )...( D  D  ...  D ) 
 11 12 1n 21 22 2n n1 n2 nn 
−7  D 
= 2  10 ln  m  in H/m
D 
 sx 

- where Dm is the Geometric Mean Distance (GMD) between all the subconductors within
conductor bundle X and all the subconductors within conductor bundle Y
(mutual geometric mean distance), in m

Dsx is the Geometric Mean Radius (GMR) of the subconductors within conductor
bundle X (self geometric mean distance)(include correction factor for
subconductor radius), in m

D 
Similarly LY' = 2  10 −7 ln  m  in H/m
 Dsy 
 

If the conductor bundles are far apart (distance D apart)

' −7  D 
LX  2  10 ln   in H/m
D 
 sx 

 D 
LY'  2  10 −7 ln   in H/m
 Dsy 
 

2.1.5 Geometric Mean Radius (GMR) of a conductor bundle

For n equally spaced subconductors of radius r ( also r = 1 ) on a circle of radius R (usual


case)

n − r
Ds = n( e 4 r )R n −1 = n n( 0,7788 r )R n −1

12 ELEN5008A – Power Systems


2.1.6 Inductance per unit length of a symmetrically spaced three-phase conductor geometry

The symmetrically spaced three-phase conductor geometry is as follows (all conductors have the
same radius, r )

With reference to the above figure, phase 1 is in position A , phase 2 is in position B and phase
3 is in position C .

The flux linkages per unit length associated with phase 1 are (requires sum of currents in all the
phases to be zero)

  1  1   1 
1 ' = 2  10 −7  I1ln   + I 2ln   + I 3ln   
  r'  D  D 

Substituting I1 + I 2 + I 3 = 0 or I 2 + I 3 = − I1

  1  1 
- we get 1 ' = 2  10 −7  I1ln   − I1ln   
  r'   D 
−7 D
= 2  10 I1ln   in Webers/m
 r' 

The self-inductance per unit length per phase is (requires sum of currents in all the phases to be
zero)
 ' −7  D 
L' = 1 = 2  10 ln   in H/m
I1  r' 

ELEN5008A - Power Systems 13


2.1.7 Inductance per unit length of an asymmetrically spaced three-phase conductor geometry

The figure below illustrates an asymmetrically spaced three-phase conductor geometry (all
conductors have the same radius, r )

Transposition will be assumed, in that phase 1 is in position A for a third of the line length, in
position C for the second third and in position B for the final third of the line length.

Phase 2 starts in position B and phase 3 starts in position C , each rotating through the
positions following phase 1 .

This ensures balanced flux linkages since, on average, each phase experiences the same magnetic
field distribution.

Transposition of the three-phase line is illustrated diagrammatically in the figure below.

The flux linkage per unit length associated with phase 1 in position A is (requires sum of
currents in the three phase conductors to be zero)

  1   1   1 
1 ' = 2  10 −7  I1ln   + I 2ln   + I 3ln  
  r'    D 
  Dab   ca  

14 ELEN5008A – Power Systems


The flux linkage per unit length associated with phase 1 in position C is (requires sum of
currents in the three phase conductors to be zero)

  1   1   1 
1 ' = 2  10 −7  I1ln   + I ln   + I ln  

  r'  2  Dca  3  Dbc  

The flux linkage per unit length associated with phase 1 in position B is (requires sum of
currents in the three phase conductors to be zero)

  1   1   1 
1 ' = 2  10 −7  I1ln   + I 2ln   + I 3ln  
  r'  D   D 
  bc   ab  

Summing these three flux linkages per unit length

−7
  1  1   1 
31 ' = 2  10  3I1ln   + I 2ln   + I 3ln  
  r'    D D D 
  Dab Dbc Dca   ab bc ca 

Substituting (this is why we need the line to be transposed – we need the second and third ln()
terms to be the same – to give a single value of self-inductance that applies to all three phase
conductors)
I1 + I 2 + I 3 = 0 or I 2 + I 3 = − I1

−7
  1   1 
- we get 31 ' = 2  10  3I1ln   − I1ln  
  r'  D D D 
  ab bc ca 

3D D D 
Or 1 ' = 2  10 −7 I1ln  ab bc ca 
in Webers/m
 r' 
 

The average self-inductance per unit length per phase is therefore

' 3D D D 
L' = 1 = 2  10 −7 ln  ab bc ca  in H/m
I1  r' 
 

If each phase comprises a conductor bundle of Geometric Mean Radius (GMR), Ds

−7 D 
L' = 2  10 ln  m  in H/m
 D 
 s 

- where Geometric Mean Distance (GMD), Dm = 3 Dab Dbc Dca

ELEN5008A - Power Systems 15


2.2 Capacitance per unit length

2.2.1 Potential difference between two points caused by a line charge, q'

Consider the following line charge geometry (charge per unit length, q' , is assumed positive)

P1 (arbitrary point in space)

D1

D2 P2 (arbitrary point in space)


q'

q'
The electric field at point P1 is EP1 = (pointing away from q' )
2 D1
q'
The electric field at point P2 is EP2 = (pointing away from q' )
2 D2

Thus the potential of point P2 with respect to point P1 is given in the equation below (voltage is
the negative of the integral of the electric field)(in the drawing above, point P2 is closer to the
line charge q' than point P1 and hence we expect the voltage between point P1 and point P2 to
be positive (point P2 is at a higher potential) – work has to be done to move a positive charge
from point P1 to point P2 )
D2 D2
q' q'  D2 
V21 = −  Ex dx = −  2 x
dx = − ln  
2  D1 
D1 D1

- where q' is the line charge per unit length, in C/m


 is the permittivity of air, in F/m

It is only the radial distances that are important (the angle has no effect on the voltages between
the two points)(equipotential circles).

2.2.2 Potential difference between two conductors in a group of conductors

Consider a group of n conductors, carrying line charges q1 ' ,q2 ', ... , qn ' , in C/m.

The radii of the conductors are r1 ,r2 , ... , rn , in m

The potential of conductor 2 with respect to conductor 1 is (need to consider the effect of the
line charges present on all the conductors)(line charges are assumed to be at the centre of the
respective conductor)(requires sum of the line charges to be zero)

16 ELEN5008A – Power Systems


1  D   r  D  D  
V21 = −  q1 ' ln  21  + q2 ' ln  2  + q3 ' ln  23  + ... + qn ' ln  2n  
2   r D  D  
  1   D12   13   1n  

The second RHS term needs special attention - the ratio within the ln() term is inverted to make
the ln() term negative. Then −1  V12 (this inverted term) = V21 .

It is assumed that the electric field within a conductor is negligible so that the integral of the
electric field is from the surface of the respective conductor (from the conductor radius).

2.2.3 Capacitance per unit length of a symmetrically spaced three-phase conductor geometry

The symmetrically spaced three-phase conductor geometry is shown below (all phase conductors
have the same radius, r )(requires sum of the line charges to be zero)

1  D  r   D 
V21 = −  q1 ' ln   + q2 ' ln   + q3 ' ln   
2   r  D  D 
1  D  r 
=−  q1 ' ln   + q2 ' ln   
2   r   D 

1  D  r 
Similarly V31 = −  q1 ' ln   + q3 ' ln   
2   r   D 

1  D  r  D  r 
Thus V21 + V31 = −  q1 ' ln   + q2 ' ln   + q1 ' ln   + q3 ' ln   
2   r  D  r   D 
1  D  r 
=−
2 
 r 
(
2q1 ' ln   + q2 ' + q3 ' ln   )
 D 

Substituting q1 ' + q2 ' + q3 ' = 0 or q2 ' + q3 ' = −q1 '

1   D 
- we get V21 + V31 = −  3q1 ' ln   
2   r 

ELEN5008A - Power Systems 17


If we consider the phasor diagram illustrating the relationship between the (assumed balanced)
line voltages and (assumed balanced) phase-neutral voltages (the neutral is not necessarily at
earth potential)

Thus V12 = V1N − V2N

Or V21 = V2N − V1N

= ( −0,5 − j0,866 )V1N − V1N

= ( −1,5 − j0,866 )V1N

Similarly V31 = ( −1,5 + j0,866 )V1N

Therefore V21 + V31 = − 3V1N

Combining this with the previous equation

V +V q ' D
V1N = − 21 31 = 1 ln  
3 2  r 

q ' 2
Thus C'1 = 1 = in F/m
V1N D
ln  
 r 

- which is the capacitance per unit length to the neutral point (the line charge is that for the
respective conductor).

Another way of looking at this is that the charge on the respective conductor is related to the
phase-neutral voltage on the conductor via the respective capacitance to the neutral point).

18 ELEN5008A – Power Systems


2.2.4 Capacitance per unit length of an asymmetrically spaced three-phase conductor geometry

The asymmetrically spaced three-phase conductor geometry is as follows (all phase conductors
have the same radius, r )(requires sum of the line charges to be zero)

Transposition will be assumed, in that phase 1 is in position A for a third of the line length, in
position C for the second third of the line length and in position B for the final third of the line
length.

Phase 2 starts in position B and phase 3 starts in position C , each rotating through the
positions following phase 1 .

This ensures a balanced system since, on average, each phase experiences the same electric field
distribution.

The potentials of phase 2 with respect to phase 1 for phase 1 in positions A , C and B are

1  D   r  D 
V21 = −  q1 ' ln  ab  + q2 ' ln   + q3 ' ln  bc 
2   r   D 
  Dab   ca 

1  D   r  D 
V21 = −  q1 ' ln  ca  + q2 ' ln   + q3 ' ln  ab 
2   r  D  D 
  ca   bc 

1  D   r  D 
V21 = −  q1 ' ln  bc  + q2 ' ln   + q3 ' ln  ca 
2   r  D  D 
  bc   ab 

Adding these together

1   Dab Dbc Dca   r3  D D D 


3V21 = −  q ' ln   + q2 ' ln   + q3 ' ln  ab bc ca 
2  1  3
  D D D 
 r  Dab Dbc Dca   ab bc ca 

ELEN5008A - Power Systems 19


1  3D D D   r 
Or V21 = − q1 ' ln  ab bc ca  + q ' ln  
2   r  2 3D D D 
    ab bc ca 

1  3D D D   r 
Also V31 = − q1 ' ln  ab bc ca  + q ' ln  
2   r  3 3D D D 
    ab bc ca 

1  3D D D   r 
Thus V21 + V31 = − 2q1 ' ln  ab bc ca  + ( q ' + q ')ln  
2   r  2 3 3D D D 
    ab bc ca 

Substituting q1 ' + q2 ' + q3 ' = 0 or q2 ' + q3 ' = −q1 '

1  3D D D 
- we get V21 + V31 = − 3q1 ' ln  ab bc ca 
2   r 
  

Previously we showed that V21 + V31 = −3V1N

Combining with the previous equation


q '  3 Dab Dbc Dca 
V1N = 1 ln  
2  r 
 

q1 ' 2 2


Thus C' = = =
VN1 3D D D  D 
ln  ab bc ca  ln  m 
 r   r 
 

- where Dm is the Geometric Mean Distance (GMD) between phases.

This is the capacitance per unit length to the neutral point (the line charge is that for the
respective conductor).

For conductor bundles replacing the phase conductors

2
C' = in F/m
 D 
ln  m 
 Deq 
 

- where Dm is the Geometric Mean Distance (GMD) between the conductor bundles
Deq is the Geometric Mean Radius (GMR) of the phase conductor bundles, ignoring the
correction to the radius of each subconductor to take the internal field into account
(we assume no internal electric field in the subconductors)

20 ELEN5008A – Power Systems


Deq = n  r  R n −1
n

2.2.5 Effect of an earth plane

To account for an earth plane, image conductors with charges − q'1 , − q'2 and − q'3 need to be
positioned beneath the earth plane in a mirror arrangement

A
Dca
Dab
C Dbc
Hcb' B
Hbc'
Hac' Hab'

Hcc' Haa' Hbb'

Hca' Hba'
Earth plane

B'
C'

A'

Both the actual charges and the image charges contribute to the potential between each of the
actual conductors and can be readily introduced into the capacitance calculation.

This is done in Grainger and Stevenson, pp 183-186.

ELEN5008A - Power Systems 21


If H aa' is the distance from position A to its image and H ab' is the distance from position A to
the image of position B , etc

2
C' = in F/m
 D   3 H H H 
ln  m  − ln  ab' bc' ca' 
 Deq   3 H H H 
   aa' bb' cc' 

In typical geometries, the phase conductors are higher above the ground than the distance
between phases.

This means that the distances from the actual to the image conductors are approximately equal
and that
 3 H H H 
ln  ab' bc' ca' 
 ln( 1) = 0
 3 H H H 
 aa' bb' cc' 

2
Thus C' = in F/m
 D 
ln  m 
 Deq 
 

- as before.

22 ELEN5008A – Power Systems


2.3 Resistance per unit length

2.3.1 Magnetic field effect on the current density

Consider the following section of a solid cylindrical conductor

In this analysis, two equations will be derived in which the magnetic field intensity, H x , is
expressed in terms of the current density, J x (now a function of radius).

The first equation is based on Ampere’s Law (the magnetic field intensity, H x , along a closed
path is equal to the enclosed current)(in this case the closed path is a circle of radius x and
circumference 2 x )
I
Hx = x
2 x

- where H x is the magnetic field intensity at radius x , in A/m


I x is the current flowing within a radius x , in A
x x x
I x =  ( area  current density ) =  ( 2 xdx )( J x ) =  2 xJ x dx
0 0 0

J x is the current density at radius x , in A/m2

x
Thus 2 xH x = I x =  2 xJ x dx
0

x
Or xH x =  xJ x dx
0

ELEN5008A - Power Systems 23


Differentiating this first equation partially with respect to x (in the second equation we derive
next we differentiate H x with respect to time)(we differentiate the LHS by parts)

H x
Hx + x = xJ x
x

H x 1
This gives the first equation + H = Jx
x x x

The second equation is obtained by considering the two axial voltage drops associated with the
loop aa' b' ba in the above figure.

The axial resistive voltage drop along line ab of length  at radius x is derived from the
v x
relationship between the electric field , resistivity  in Ωm and current density J x (known

as the 3D Ohm’s law)
 v x =  J x

The axial resistive voltage drop along line a' b' of length  at radius x + dx is similarly

 J 
vx + dx =   J x + x dx 
 x 

The net axial resistive voltage drop around the loop aa' b' ba is the difference between these two
axial resistive voltage drops (the voltage drop along the length aa' and bb' is asymptotically
zero)
J x
vx + dx − vx =  dx
x

In addition, the emf induced around the loop aa' b' ba can be obtained from the flux passing
through the loop (flux linkage is equal to the enclosed flux because there is only one turn)

 = Bx  dx

=  H x  dx

 H x
The induced emf is (Faraday’s law) e = − = −  dx
t t

24 ELEN5008A – Power Systems


Equating this to the net axial resistive voltage drop and assuming sinusoidal conditions (replace
H x with H xe jt and J x with J xe jt )

( H x e jt ) ( J x e jt )
−  dx =  dx
t x

J x jt
Or − j H x e jt dx =  e dx
x

 J x
This gives the second equation Hx = j
 x

H x 1
But from the first equation + H = Jx
x x x

2
  J x 1  J x
Substituting for H x j + j = Jx
 x 2 x  x

This gives the required equation for the current density at radius x , J x

2 J x 1 J x j
+ + J =0
x 2 x x  x

Since the time factor is no longer in the equation, the partial derivatives can be replaced by
ordinary derivatives.

In addition, the conductivity  is frequently used instead of the resistivity  , where

1
=

d 2Jx 1 dJ x
Thus + + j J x = 0
dx 2 x dx

ELEN5008A - Power Systems 25


The solution to this equation may be found by noting that it has the form of a Bessel equation of
zero order
2
d y 1 dy
+ + k2 y = 0
dx 2 x dx

The solution to this known equation form is the following power series

 ( kx )
2
( kx )
4
( kx )
6
( kx )
8
( kx )
10 
y = a0  1 − + − + − + ... 
 22 224 2 224 26 2 224 26 28 2 224 26 28 210 2 
 

Thus the solution to the previous equation is obtained by substituting

k 2 = j = jm2

- where m= 

 ( mx )
2
( mx )
4
( mx )
6
( mx )
8
( mx )
10 
Thus J x = a0  1 − j − + j + − j − ... 
 22 224 2 224 26 2 224 26 28 2 2 24 26 28 210 2 
 
 ( mx )
4
( mx )
8   ( mx )2 ( mx )
6
( mx )
10 
= a0  1 − + − ...  + ja0  − + − + ... 
 224 2 224 26 28 2   22 2 24 26 2 2 24 26 28 210 2 
   
= a0 ( ber( mx ) + j bei( mx ))

4 8
( mx ) ( mx )
where ber( mx ) = 1 − + − ...
224 2 224 26 28 2

2 6 10
( mx ) ( mx ) ( mx )
bei( mx ) = − + − + ...
22 224 26 2 224 26 28 210 2

The constant a0 may be evaluated relative to the current density at the conductor surface, J r
(setting x = r )
J r = a0 ( ber( mr ) + j bei( mr ))

ber( mx ) + j bei( mx )
Thus Jx = J
ber( mr ) + j bei( mr ) r

26 ELEN5008A – Power Systems


2.3.2 Internal impedance per unit length of a solid cylindrical conductor, Z i '

Considering only the internal impedance of a conductor (ignoring the external magnetic flux)(we
are focussing only on skin effect), the axial voltage drop per unit length along the surface layer is
purely resistive (there is no external magnetic flux linking it and hence no inductance) – this
equation seems simple but we do not as yet know the surface current density, J r

J
Vi ' = r

 J x 1 J x 1  J x 
From the second equation Hx = j = j = j  
 x  x m 2  x 

1  J x 
At the surface ( x = r ) Hr = j  
m2  x x=r
d 
 dx ( ber( mx ) + j bei( mx )) 
= j x=r Jr
m2 ( ber( mr ) + j bei( mr ))
m( ber'( mr ) + j bei'( mr ))
=− j Jr
m2 ( ber( mr ) + j bei( mr ))
bei'( mr ) − j ber'( mr )
= J
m( ber( mr ) + j bei( mr )) r

2 r bei'( mr ) − j ber'( mr )
But I = 2 rH r = J
m ber( mr ) + j bei( mr ) r

This is the equation we wanted - to calculate the surface current density from the total current.

V' J m ber( mr ) + j bei( mr )


Therefore Zi ' = i = r =
I  I 2 r bei'( mr ) − j ber'( mr )

This equation can be split into real (resistance) and imaginary (internal inductive reactance)
components – while Vi ' is in phase with J r (by definition), the total current, I , is not.

ELEN5008A - Power Systems 27


The resistance per unit length is (real part of Z i ' )

m  ber( mr )bei'( mr ) − bei( mr )ber'( mr ) 


Ri ' =   in /m
2 r  ( bei'( mr ))2 + ( ber'( mr ))2 

1 1
But RDC ' = R0 ' = = in /m
conductance  area  r 2

Ri ' mr  ber( mr )bei'( mr ) − bei( mr )ber'( mr ) 


Thus =  
R0 ' 2  ( bei'( mr ))2 + ( ber'( mr ))2 

The internal inductive reactance per unit length is (imaginary part of Z i ' )

m  bei( mr )bei'( mr ) + ber( mr )ber'( mr ) 


 Li ' =   in /m
2 r  ( bei'( mr ))2 + ( ber'( mr ))2 

m  bei( mr )bei'( mr ) + ber( mr )ber'( mr ) 


Or Li ' =   in /m
2 r  ( bei'( mr ))2 + ( ber'( mr ))2 

But from the previous equation for the internal inductance per unit length where skin effect was
not present
 m2
LDC ' = L0 ' = =
8 8

Li ' 4  bei( mr )bei'( mr ) + ber( mr )ber'( mr ) 


Thus =  
L0 ' mr  ( bei'( mr ))2 + ( ber'( mr ))2 

28 ELEN5008A – Power Systems


These equations are usually presented in the form of curves

1,7

1,6

1,5

1,4
R' i
R'0
1,3

1,2

1,1

1,0
0 0,5 1,0 1,5 2,0 2,5 3,0 3,5 4,0

mr
m = 

Skin Effect: Resistance Ratio
 = r o

r = radius, in m

1
The resistance per unit length then becomes ks  , where k s is the above resistance ratio.
A

ELEN5008A - Power Systems 29


1,0

0,95

0,9

0,85
L'i
L'0
0,8

0,75

0,7

0,65
0 0,5 1,0 1,5 2,0 2,5 3,0 3,5 4,0

mr
m = 

Skin Effect: Internal inductance Ratio
 = r o

r = radius, in m


− ks r
The Geometric Mean Radius (GMR) of the single conductor then becomes r' = r( e 4 )

where k s is the above internal inductance ratio.

A reminder that skin effect only influences the inductance associated with the internal flux.

30 ELEN5008A – Power Systems


CHAPTER 3: STEADY-STATE ANALYSIS OF
TRANSMISSION LINES

3.1 Equivalent circuit of a transmission line

The equivalent circuit of a transmission line is based on the resistance, inductance, capacitance
and conductance per unit length of the line – as derived in the previous chapter.

For an overhead line in a balanced three-phase configuration or having transposition (phase


conductor positions regularly swapped), the parameters per phase per meter are

 D 
L' = o ln  m  in H/m
2  Ds 

2 o
C' = in F/m
 D 
ln  m 
 Deq 
 


R' = ks in /m
A

- where Dm is the Geometric Mean Distance between phase conductors, in m


Ds is the Geometric Mean Radius for each phase conductor, in m
Deq is the Geometric Mean Radius for each phase conductor assuming no internal
electrical field within the subconductors, in m
k s is the resistance skin-effect correction factor
 is the resistivity of the phase conductor material, in Ωm
A is the total cross-sectional area per phase conductor, in m2

3.1.1 Short lines (< 50 km)

For short lines, it is only necessary to use the series components to give a good approximation of
the line behaviour ( is the line length, in m)

R = R'
X L =  L'

ELEN5008A - Power Systems 31


3.1.2 Medium length lines (50 km – 200 km)

For medium length lines the ‘nominal-π model’ or the ‘nominal-T model’ is used (shunt
capacitance and shunt conductance become significant)

where Z = ( R' + j L')

( G' + jC')
Y =
2

( R' + j L')
ZT =
2

YT = ( G' + jC')

3.1.3 Long lines (> 200 km)

For long lines, a more precise (equivalent) model is derived from the hyperbolic equations which
describe the exact steady-state-sinusoidal behaviour of any transmission line

VS = VR cosh(  ) + I R Z o sinh(  )

V
I S = I R cosh(  ) + R sinh(  )
Zo

- where  is the propagation constant of the transmission line (complex constant)

 = ( R' + j L')( G' + jC')

Z o is the characteristic impedance of the transmission line (complex constant)

R' + j L'
Zo =
G' + jC'

32 ELEN5008A – Power Systems


These equations can be used to derive the component values in the ‘equivalent-π model’ or
‘equivalent-T model’ for long lines (same circuits as in Section 3.1.2 - except replace in name
‘nominal’ with ‘equivalent’).

The voltage equation for the ‘equivalent-π model’ is derived as follows (phasor equation)

VS = VR + ( I R + VRY )Z

= VR ( 1 + Y Z ) + I R Z

- where VS and VR are the phase-neutral voltage phasors at the sending end and receiving end
respectively
I R is the receiving end current phasor

Comparing with the hyperbolic voltage equation for long lines

VS = VR cosh(  ) + I R Z o sinh(  )

This implies Z = Z o sinh(  )

1 + Y Z = cosh(  )

cosh(  ) − 1 cosh(  ) − 1
Y = =
Z Zo sinh(  )

ELEN5008A - Power Systems 33


Similarly for the ‘equivalent-T model’

I S = I R + VT YT

I S = I R + (VR + I R ZT )YT

= I R ( 1 + ZT YT ) + VRYT

Comparing with the hyperbolic current equation for long lines

V
I S = I R cosh(  ) + R sinh(  )
Zo

sinh(  )
This implies YT =
Zo

1 + ZT YT = cosh(  )

cosh(  ) − 1  cosh(  ) − 1 
ZT = =  Zo
YT  sinh(  ) 

- all we need to know are the values for  , Z o and .

Typically, the voltages at the sending-end and at the receiving-end as well as the power flow
along the line are of interest - a simple method of analysis is described in the next section.

34 ELEN5008A – Power Systems


3.2 The receiving-end power diagram (so-called because the receiving-end active-power and
reactive power are specified)

The phase-neutral voltage phasor (single phase) equation for the line must be expressed as

VS = VR ( 1 + Y Z ) + I R Z = AVR + BI R

- where VS is the sending-end phase-neutral voltage phasor, in V


VR is the receiving-end phase-neutral voltage phasor, in V
I R is the receiving-end current phasor, in A
A ( = A  ) and B ( = B  ) are complex constants, determined from the
equivalent circuit or the hyperbolic equations (assumed known from the line
geometry)

For a short line (< 50 km) A = 1 + Y Z 10 ( since Y 0)

B = Z = R + jX L

For a medium length line A = 1 + Y Z

B = Z = R + jX L

For a long line (> 200 km) A = cosh(  )

B = Z o sinh(  )

The receiving-end complex power per phase is given by ( VR is the reference phase-neutral
voltage phasor – phase angle of zero)


SR = VR I R = VR I R (  )

Note that with this definition, a positive value of  indicates that the load current phasor lags the
load voltage phasor – a lagging load power factor.

ELEN5008A - Power Systems 35


Also the previous equation, but now expressed in terms of the complex conjugates of the voltage
phasors and current phasor, can be expressed as ( VS is the sending-end phase-neutral voltage
phasor)
VS = AVR + BI R

V
Multiplying by R to get a phasor equation in terms of the receiving-end voltages and currents
B
per phase (single phase) complex power VR I R

VSVR AVRVR 
= + VR I R
 
B B

The above equation can also be written in polar form ( VR is the reference phase-neutral voltage
phasor – has a phase angle of zero)

2
VS VR ( − ) A ( − ) VR
= + VR I R (  )
B ( −  ) B ( −  )

2
VS VR A VR
Or (  −  ) = (  −  ) + VR I R (  )
B B

- where A and B are assumed known


 is the angle associated with the complex parameter A (known)
 is the angle associated with the complex parameter B (known)
 is the load angle (difference between the phase angle of the phase-neutral voltage
phasor VS and the phase angle of the phase-neutral voltage phasor VR ), in rads
(do not confuse with the load impedance angle,  )
 is the load impedance angle, in rads

36 ELEN5008A – Power Systems


The receiving-end power diagram (receiving-end voltages and currents specified) is therefore as
follows (the above single-phase equation gives the per-phase powers – the three-phase powers
can be obtained by multiplying each of the terms by three or by replacing the phase-neutral
voltages with the line voltages in the product-of-two-voltages-terms and replacing the
VR I R (  ) term with 3 VR I R (  ) )

Receiving-end VARs

3 VR IR
Lagging VARs


Receiving-end Watts

2
-( - ) VS VR
A VR B
Leading VARs
B
 -  (a function of the load angle, )

 -  (not a function of the load angle, )

 -  (a function of the load angle, )

All voltages now line voltages

All powers (MWs and MVARs) now three-phase values

2
A VR VS VR
An inductive load reduces VR relative to VS and hence is smaller than as
B B
seen above (assuming A 1 ).

While  and  have been previously determined and hence  −  can be calculated,  − 
can be calculated using simple trigonometry

 A VR
2 
 3 VR I R cos  + cos(  −  ) 
 B 
 −  = cos −1  
 VS VR 
 
 B 
 

Also  = −(  −  ) + 

ELEN5008A - Power Systems 37


Making VR = VS is achieved by adding capacitance (phase-to-earth) at the load end as shown
below

Receiving-end VARs

3 VR IR

Reactive power (capacitive)



that must be added at load end
to make VR = VS
2
-( - )
A VR
B

-

- VS VR
B
-

If the load is disconnected (the transmission line is left open-circuited at the load end), VR rises
2
A VR VS VR
relative to VS due to the line capacitance and hence is larger than as seen
B B
below (assuming A 1 ).

This is remedied by adding inductance (phase-to-earth) at the load end as shown below (known
as a line (shunt) reactor)- especially important if the transmission is energised at the sending-end
without load connected at the receiving-end).

2
Equivalent reactive power Reactive power (inductive)
A VR (capacitive) at load end that must be added at load
that causes VR > VS end to make VR = VS
B

VS VR
B

38 ELEN5008A – Power Systems


3.3 Calculating the power limit from the power diagram

From the power diagram, it is possible to derive an expression for the maximum power that a
transmission line can deliver to a load.

The load real power is the projection of VR I R onto the real power (horizontal) axis

Load real power = VR I R cos 

2
VS VR A VR
= cos(  −  ) − cos(  −  )
B B

In typical transmission systems, the series impedance R + j L is dominated by j L and hence


  90 .

In addition, the shunt conductance of a transmission line is negligible in comparison with the
capacitive susceptance and hence   0 .

Therefore cos(  −  )  cos( 90 ) = 0 and the above equation reduces to

VS VR
Load real power = cos(  −  )
B

A further simplification follows because   90 and cos( 90 −  ) = sin(  ) .

VS VR
Thus Load real power = sin 
B

The power limit PL occurs when  = 90

VS VR
PL =
B

VS VR

Xs

VS VR
=
L

ELEN5008A - Power Systems 39


3.4 Capacitive shunt and series compensation

Shunt compensation (capacitance connected phase-to-earth) at the receiving-end changes the


effective load power factor.

This usually means that the load power factor is improved (brought closer to unity) and, as a
result, the load current is reduced.

The voltage drop along the line decreases and the load voltage increases.

With the decrease in load current the line I 2 R losses also decrease.

Shunt compensation (capacitance connected phase-to-earth) has no significant effect on the


power limit of a line apart from controlling the terminal voltages.

Series compensation (capacitance connected in series with the line) decreases the effective series
impedance of the line.

The line voltage drop is decreased and variations in load voltage with varying load decrease.

The decrease in series impedance results in an increase in the power limit of the line.

Series compensation (capacitance connected in series with the line) does not reduce the load
current and therefore does not influence the line losses.

A serious problem with series compensation is a phenomenon called subsynchronous resonance


(SSR).

The series capacitance may resonate with the inductance of the line, the inductance of the
transformers and the synchronous inductance of a generator at frequencies below the power
frequency.

These resonant frequencies may coincide with the mechanical resonant frequencies of the
turbine-generator shafts - such resonances have caused damage to turbine-generator shafts.

3.5 Voltage control using transformers with tap changers

A transformer with taps on one of its windings can be used for voltage control.

As the load increases and the load voltage drops, the transformer turns ratio is changed (via
changing the tap position along the winding), perhaps automatically, to boost the load voltage.

The changing of the transformer turns ratio (via changing the tap position along the winding)
changes both the transformer impedance and the impedances reflected through the transformer.

This must be kept in mind in analysis of a circuit containing a tap-changing transformer.

The most frequent cause of failure of large transformers is the on-load tap-changer.

40 ELEN5008A – Power Systems


CHAPTER 4: SOLVING FOR POWER FLOW / LOAD FLOW

4.1 Mesh current method and the concept of the connection matrices

Consider the simple balanced three-phase power system below

Transmission line

Generator a Generator b

Loads Loads

One phase of the above can be represented as follows

z cc ic
ia id ie ib

z aa z bb
z dd z ee
I1 I2 I3
ea eb

Neutral

- where zdd and zee are the impedances of the passive (no voltage sources) loads at the busbars
together with the shunt impedances of the nominal-π circuit representing the transmission line.

All mutual impedances between the branches are assumed to be zero (a current in one branch
(e.g. ic ) does not directly induce a voltage in another branch (e.g. branch consisting of
impedance zdd )).

The branch currents, the branch driving voltages (emfs) and the branch impedances are shown in
lower case, i.e. ia , ea and zaa in order to distinguish them from the mesh currents I 1 , I 2 and
I 3 , the mesh driving voltages E1 , E2 and E3 , the mesh self-impedances Z11 , Z 22 and Z 33 .

All currents and voltages are the rms values of sinusoidally varying quantities (50 Hz).

By convention the emfs on the left-hand side of the equals sign are positive if the respective
mesh current direction indicates power flow out of the respective emf.

Similarly, the impedance voltage drops on the right-hand side of the equals sign are positive for
impedances through which the respective mesh current flows. However, if the impedance is
shared with an adjacent mesh this impedance voltage drop is reduced by the impedance voltage
drop due to the current in the adjacent mesh (equivalent to this impedance voltage drop being
negative).

ELEN5008A - Power Systems 41


For the three meshes above, the following three mesh equations can be written relating the
branch driving voltages (emfs) and the product of the branch impedances and the mesh currents

ea = ( zaa + zdd )I1 − zdd I 2

0 = − zdd I1 + ( zcc + zdd + zee )I 2 − zee I 3

−eb = − zee I 2 + ( zbb + zee )I 3

We can also define the mesh driving voltages as the sum of the branch driving voltages along the
path the mesh current flows (if the branch driving voltage is in the opposite direction to the mesh
current then it is represented as being negative).

 E1   ea 
For the above network the mesh driving voltage vector is  E  =  E2  =  0 
 −e 
 E3   b 
 I1 
Similarly, the mesh current vector is  I  =  I 2 
 I3 

Thus

 E1   zaa + zdd − zdd 0   I   Z11 Z12


1 Z13 
 E  =  E2  =  − zdd zcc + zdd + zee − zee   I 2  =  Z 21 Z 22
  
Z 23  [ I ] = [ Z ][ I ]

 E3   0 − zee zbb + zee   I 3   Z 31 Z 32 Z 33 

Or E1 = Z11 I1 + Z12 I 2 + Z13 I 3


E2 = Z 21I1 + Z 22 I 2 + Z 23 I 3
E3 = Z 31 I1 + Z 32 I 2 + Z 33 I 3

In general, for mesh n

En = Z n1I1 + Z n2 I 2 + ... + Z nn I n + Z n( n +1 )I( n +1 ) + ...

- where Z nn (a diagonal term in the mesh impedance matrix) is the mesh self-impedance –
which is the total impedance round mesh n

Z nm (an off-diagonal term in the mesh impedance matrix) is the shared impedance
between meshes n and m (for the case of zero mutual coupling between branches)

42 ELEN5008A – Power Systems


For electrical networks, Z mn = Z nm and the off-diagonal impedances in the mesh impedance
matrix [ Z ] are symmetrical about the diagonal impedances.

The off-diagonal terms (mesh mutual impedances) are all negative when this cyclic current
convention is used.

The relationships between the branch currents and the mesh currents in the previous network are

ia = I1
ib = − I 3
ic = I 2
id = I 2 − I1
ie = I 3 − I 2

For mesh analysis, these current constraints are expressed in terms of a connection matrix [ C ]

[ i ] = [ C ][ I ]

For our original network

 ia   1 0 0 
 i   0 0 −1   I 
 ib  =  0 1 0   I 1 
 c   2
 id   −1 1 0   I 3 
 ie   0 −1 1 

1 0 0 
 0 0 −1 
And hence [C ] =  0 1 0 
 −1 1 0 
 0 −1 1 

Power invariance in the transformed equations requires that the mesh driving voltage vector
[ E ] (sum of the branch driving voltages round mesh) is related to the branch driving voltage
vector [ e ] (sum of the voltage sources within branch) by (for sinusoidal voltages and currents)

p = [ e ]t [ i*] = [ e ]t [ C*][ I*] =  [ C*]t [ e ]  [ I*] = [ E ]t [ I*]


t

- where [ C*] is the complex conjugate of [ C ] and [ C*]t is the transpose of [ C*] - note that
[ C ] is not a square matrix and hence does not have an inverse.

Thus [ E ] = [ C*]t [ e ]

ELEN5008A - Power Systems 43


And since [ C ] contains only real terms (-1 or 0 or 1)

[ E ] = [ C ]t [ e ]

We now introduce the concept of primitive circuits – decomposing the actual network into the
most basic (primitive) circuits.

The relationship between the primitive branch driving voltage vector [ e prim ] and the primitive
branch current vector [ i prim ] formed when the branches are all in their respective locations in
space but are electrically isolated and short-circuited is (for our original network)

e c(prim)
i d(prim) (=0) i e(prim)
i a(prim) z cc i b(prim)
(=0) (=0)

z aa z dd z ee z bb
i c(prim)
(=0)

e a(prim) e d(prim) e e(prim) e b(prim)


(=0) (=0)

For the primitive system [ e prim ] = [ z prim ][ i prim ]

- where (we can now see the advantage - working with diagonal impedances matrices)

 zaa 0 0 0 0 
 0 zbb 0 0 0 
[ z prim ] =  0 0 zcc 0 0 

 0 0 0 zdd 0 
 0 0 0 0 zee 

We now want to introduce the transformation between the original branch currents and branch
driving voltages and the primitive branch currents and branch driving voltages via the
transformation matrix [ A ] . The transformation matrix is the same for the currents and voltages
since the impedances remain unchanged.

Because there are as many original branches as primitive branches, [ A ] is a square matrix and
hence in this case has an inverse – compared with [ C ] which is not a square matrix and hence
does not have an inverse.
[ i ] = [ A][ i prim ]
[ e] = [ A][ e prim ]

For power invariance p = [ e ]t [ i*] = [ e prim ]t [ A]t [ A][ i prim*] = [ e prim ]t [ i prim*]

44 ELEN5008A – Power Systems


Which implies that [ A ] must have the property [ A]t = [ A] −1

The equation relating the primitive network equations and the original network mesh equations
(satisfying power invariance) is

[ E ] = [ C ]t [ e ] = [ C ]t [ A][ e prim ] = [ C ]t [ A][ z prim ][ i prim ]

= [ C ]t [ A][ z prim ][ A] −1 [ i ] = [ C ]t [ A][ z prim ][ A] −1 [ C ][ I ]

= [ Z ][ I ]

Thus [ Z ] = [ C ]t [ A][ z prim ][ A] −1 [ C ] = [ C ]t [ z prim ][ C ]

- where [ Z ] is the mesh impedance matrix for the original network.

For our original network (remember when multiplying matrices, the number of rows must equal
the number of columns)

 zaa 0 0 0 0  1 0 0
 
 0   0 0 −1 
 1 0 0 −1 0   0 zbb 0 0
[ Z ] =  0 0 1 1 −1   0 0 zcc 0 0  0 1 0 

 0 −1 0 0 1   0 0 0 zdd 0   −1 1 0 
 0 0 0 0 zee   0 −1 1 

 zaa 0 0 
 − zbb 
 1 0 0 −1 0   0 0
=  0 0 1 1 −1   0 zcc 0 

 0 −1 0 0 1   − z zdd 0 
dd
 0 − zee zee 

 zaa + zdd − zdd 0 


=  − zdd zcc + zdd + zee − zee 
 
 0 − zee zbb + zee 

- as derived initially from the basic mesh equations.

We have now derived the relationship between [ Z ] and [ z prim ] - used for determining [ Z ]
when we know [ z prim ] and the connection matrix [C] for our original network.

ELEN5008A - Power Systems 45


4.2 Nodal voltage method and connection matrices

Nodal voltage analysis is the more commonly used alternative to mesh current analysis for
solving power system network equations.

We again assume our original network

Transmission line

Generator a Generator b

Loads Loads

1
We replace the impedances with admittances (branch admittance = ) , replace
branch impedance
the voltage sources with current sources (Norton equivalent) and number the nodes as follows (0,
1 and 2 for our example network)

y cc
ia 1 ic 2 ib
id ie

I 1 = yaa e a yaa y dd y ee ybb I 2 = ybb e b

Neutral
0

The convention is that the net current on the left-hand side of the equals sign is zero and on the
right-hand side of the equals sign, currents flowing away from the node are positive.

For node 1 0 = − ia − id + ic

For node 2 0 = − ic − ie − ib

Writing these equations in terms of the branch admittances (lower case y ), branch driving
voltages (lower case e ) and the potentials of one node with respect to another node (capital V )

For node 1(KCL) 0 = yaa (V10 − ea ) + ydd (V10 ) + ycc (V12 )

For node 2 (KCL) 0 = ycc (V21 ) + yee (V20 ) + ybb (V20 − eb )

46 ELEN5008A – Power Systems


Rearranging terms and substituting V12 = V10 − V20

V21 = V20 − V10

- so that only potentials with respect to the reference node 0 remain and emfs are on the left-
hand side of the equals sign

yaa ea = ( yaa + ycc + ydd )V10 − yccV20

ybb eb = − yccV10 + ( ybb + ycc + yee )V20

 I   y + ycc + ydd − ycc   V10   Y11 Y12 


Or [ I ] =  1  =  aa = [V ] = [Y ][V ]
 I2   − ycc ybb + ycc + yee   V20   Y21 Y22 

For node 1 (using alternative name for I 1 ) ( ye )1 = I1 = Y11V10 + Y12V20

For node 2 (using alternative name for I 2 ) ( ye )2 = I 2 = Y21V10 + Y22V20

In general ( ye )n = I n = Yn1V10 + Yn2V20 + ... + YnnVn0 + Yn( n +1 )V( n +1 )0 + ...

Ynn (a diagonal term in the nodal admittance matrix) is the nodal self-admittance at node n and
is equal to the sum of all admittances connected directly to that node (includes both admittances
between node n and any other node and admittances between node n and the reference node 0 ).

In the above example Y11 = yaa + ycc + ydd

Ynm (an off-diagonal term in the nodal admittance matrix) is the admittance between nodes n
and m and is equal to the sum of all admittances connected directly between these two nodes.

In the above example (all off-diagonal terms are negative)

Y12 = − ycc

For an electrical network Ynm = Ymn

For the node numbered n , ( ye )n is the sum of the products of the branch driving voltages and
the admittances of the branches containing them which are connected directly to node n .

It is effectively the current injected into the network at that node.

This nodal current is positive, i.e. it is directed towards that node, as occurs in both current
sources of the previous network.

ELEN5008A - Power Systems 47


If the circuit contained a third node with a source of impedance znn having an emf en directed
away from a node 3, then the current injected into node 3 would be I 3 = − ( ynnen ) .

As another example, consider the following 5-bus network

1 2 3 4

The equivalent circuit showing all admittances is as follows (injected currents also included)

I2 I3
1 2 y23 3 4
y T1 y T2
I1 I4
y23s y23s
y25s y35s
y25 y35
y25s y35s

I5

The nodal admittance matrix equation is

 I1   Y11 Y12 Y13 Y14 Y15   V1 


 I  Y Y22 Y23 Y24 Y25   V2 
 I 2  =  Y21 Y32 Y33 Y34 Y35   V3 
 3   31  
 I 4   Y41 Y42 Y43 Y44 Y45   V4 
 I 5   Y51 Y52 Y53 Y54 Y55   V5 

The terms in the 5  5 nodal admittance matrix are

Y11 = yT1 ; Y12 = − yT 1 ; Y13 = Y14 = Y15 = 0

Y21 = − yT 1 ; Y22 = ( yT 1 + y25 + y25s + y23 + y23s ) ; Y23 = − y23 ; Y24 = 0 ; Y25 = − y25

Y31 = 0 ; Y32 = − y23 ; Y33 = ( yT 2 + y35 + y35s + y23 + y23s ) ; Y34 = − yT 2 ; Y35 = − y35

Y41 = Y42 = 0 ; Y43 = − yT 2 ; Y44 = yT 2 ; Y45 = 0

Y51 = 0 ; Y52 = − y25 ; Y53 = − y35 ; Y54 = 0 ; Y55 = ( y35 + y35s + y25 + y25s )

48 ELEN5008A – Power Systems


Going back to the original network at the start this chapter, the relationships between the branch
voltages (do not confuse with the branch driving voltages (emfs)) and the nodal voltages are

va = V10

vb = V20

vc = V20 − V10

vd = V10

ve = V20

These voltage constraints may be expressed using a connection matrix [ C ] where

 va   1 0
v  0 1  V 
[ v ] =  vc  =  −1
b
  1   10  = [ C ][V ]
 vd   1 0   V20 
 ve   0 1 

1 0
0 1
For our original network [ C ] =  −1 1
1 0
 0 1 

Power invariance in the transformed equations requires the following relationship between the
nodal currents and the branch currents (note that we must use the branch driving voltages (emfs))

p = [ i ]t [ e*] = [ i ]t [ C*][ E*] =  [ C*]t [ i ]  [ E*] = [ I ]t [ E*]


t

Thus [ I ] = [ C*]t [ i ]

And since [ C ] contains only real terms [ I ] = [ C ]t [ i ]

The relationship between the branch current matrix [ i ] and the branch voltage matrix [ v ] can
be obtained using Norton’s theorem.

Thus, all branches in the network may be shown separately as their admittances, each active one
(containing a voltage source) being shunted by a current source ( ye ) and the branch currents
must each be given by the product of the branch admittance and the branch voltage so that

[ i ] = [ y ][ v ]

ELEN5008A - Power Systems 49


Once again we introduce the concept of primitive circuits – decomposing the actual network into
the most basic (primitive) circuits.

For our original network

y cc

vc(prim)
(=0)
va(prim) yaa vd(prim) ydd yee ve(prim) ybb vb(prim)
(=0) (=0)

ycc e c(prim)
yaa e a(prim) ydd e d(prim) yee e e(prim) ybbe b(prim)
(=0)
(=0) (=0)

For the primitive circuits the relationship between the branch current sources and the branch
voltages is
[ i prim ] = [ y prim ][ v prim ]

 yaa 0 0 0 0 
 0 ybb 0 0 0 
- where [ y primitive ] =  0 0 ycc 0 0 

 0 0 0 ydd 0 
 0 0 0 0 yee 

If we now want to include the transformation between the original branch currents and branch
driving voltages and the primitive branch currents and branch voltages via the transformation
matrix [ B ] – because there are as many original branches as primitive branches [ B ] is a
square matrix and hence in this case has an inverse – compared with [ C ] which is not a square
matrix and hence does not have an inverse.

The transformation matrix is the same for the currents and voltages because the admittances
remain unchanged.
[ i ] = [ B ][ i prim ]
[ v ] = [ B ][ v prim ]

For power invariance p = [ v ]t [ i*] = [ v prim ]t [ B ]t [ B ][ i prim*] = [ v prim ]t [ i prim*]

Which implies that [ B ] must have the property [ B ]t = [ B ] −1

Thus [ I ] = [ C ]t [ i ] = [ C ]t [ B ][ i prim ] = [ C ]t [ B ][ y prim ][ v prim ]

50 ELEN5008A – Power Systems


= [ C ]t [ B ][ y prim ][ B ] −1 [ v ] = [ C ]t [ B ][ y prim ][ B ] −1 [ C ][V ]

= [Y ][V ]

Thus [Y ] = [ C ]t [ B ][ y prim ][ B ] −1 [ C ] = [ C ]t [ y prim ][ C ]

- where [ Y ] is the nodal admittance matrix for the original interconnected network.

For our original network (remember when multiplying matrices the number of rows of the RHS
matrix must equal the number of columns of the LHS matrix)

 yaa 0 0 0 0  1
 0
 0 ybb 0 0 0  0 1
[Y ] =  1 0 −1 1 0   0 0 ycc 0 0   −1 1
 0 1 1 0 1   
 0 0 0 ydd 0  1 0
 0 0 0 0 yee   0 1 

 yaa 0 
 0 ybb 
=  1 0 −1 1 0  −y ycc 
 0 1 1 0 1   y cc 0 

 dd
 0 yee 

 y + y + ydd − ycc 
=  aa cc
 − ycc ybb + ycc + yee 

- as derived initially from the basic nodal equations.

We have now derived the relationship between [ Y ] and [ y prim ] - used for determining [ Y ]
when we know [y prim ] and the connection matrix [C] .

ELEN5008A - Power Systems 51


4.3 Application of the nodal voltage (admittance) method to the solution of power system
power flow / load flow problems

Power flow analysis involves the determination of the network voltages and the network currents
under steady-state (50Hz) conditions, to investigate whether
- all voltages are within specified limits
- network components (transmission lines, cables and transformers) are not overloaded
especially under contingency conditions (other transmission lines, cables and transformers
disconnected due to faults or general maintenance)
- excessive reactive power is not being transferred across transmission lines, cables and
transformers – prefer to generate reactive power (e.g. capacitors) where it is consumed (e.g.
at an inductive load)
- the operating conditions following proposed extensions to the network (additional loads,
generation, transmission lines, cables and transformers)

Power flow analysis provides as output


- the magnitude and phase angle of the voltages at all the busbars
- the active power flow along all the transmission lines and cables and through the
transformers
- the reactive power flow along all the transmission lines and cables and through the
transformers
- the losses in the network

The establishment of a floating bus – infinite busbar – is required because the power input at (at
least) one busbar is unknown until the power flow study is done because the total power loss in
the system can only then be determined.

The first step of the power flow analysis is to determine the nodal admittance matrix [ Y ] using
- the values for the individual impedances (admittances)
- the way the individual impedances (admittances) are interconnected

The advantages of the nodal voltage (admittance) method include


- network components (generators and loads) are inherently connected in parallel
- the number of equations is inherently less
- there is no need to combine parallel branches before forming equations (has no effect on the
number of equations)(huge advantage)
- there is no difficulty with cross-over branches (huge advantage)
- transformers with tap-changers can be easily represented

The disadvantages of the nodal voltage (admittance) method include


- converting all impedances to admittances
- nodal voltages must be calculated to a high degree of accuracy as the branch currents depend
on small differences between them

52 ELEN5008A – Power Systems


4.3.1 Power flow / load flow equations (nodal equations)

The starting point for power flow / load flow analysis is a single-line diagram of the power
system including
- busbar (node) data (two of the following four parameter values must be already known – the
remaining two are then solved)
- voltage phasor magnitude Vk at each busbar
- voltage phasor phase angle  k at each busbar
- for each busbar the active power generated by local generators (MW) less the active
power consumed by local loads (MW rather than represented by a shunt conductance)
Pk
- for each busbar the reactive power generated by local generators (MVAr) less the
reactive power consumed by local loads (MVAr rather than represented by a shunt
susceptance) Qk
- transmission line and cable impedances
- transformer impedances

The net powers supplied to the busbar k are separated into generator and load terms

Pk = PGk − PLk
Qk = QGk − QLk

Note that when busbar k is a load bus with no generators, Pk is negative (local load represented
as injecting negative active power into the busbar).

Similarly if the load is inductive with no local reactive power sources, Qk is negative (local load
represented as injecting a negative reactive power into the busbar)

To other busbars

Vk  k Busbar, k

PGk Q Gk PLk Q Lk

Gen Load

ELEN5008A - Power Systems 53


For power flow / load flow analysis, each busbar (node) is categorized into one of the following
three busbar types
- Floating busbar (slack busbar or swing busbar or infinite busbar) - numbered 1 – nodal
voltage phasor magnitude and phase angle predefined for this single busbar (typically
V1 1 = 1,00 in pu) - P1 and Q1 are then solved for this single busbar
- Load busbars (PQ busbars) – Pk and Qk are already known for these busbars - the nodal
voltage phasor magnitude and phase angle ( Vk and  k ) are then solved for this type of
busbar (most busbars are of this type – being busbars to which generators (not controlling
the busbar voltage) and loads (not controlling the busbar voltage) are connected)
- Voltage controlled busbars (PV busbars) – Pk and Vk are already known for these busbars
- the nodal voltage phase angle and reactive power injection (  k and Qk ) are then solved
for these busbars (examples of these busbars are where generators are located (injecting a
particular MW and controlling the busbar voltage via injection of reactive power), shunt
capacitors are located (injecting zero MW and controlling the busbar voltage via injection of
reactive power) and static var compensators (SVCs) are located (injecting zero MW and
controlling the busbar voltage via injection of reactive power))

Transmission lines, cables and transformers are represented by their nominal −  (or equivalent-
π) circuits.

The admittance matrix [Y ] can be constructed from transmission line, cable and transformer
impedance data.

The elements of [Y ] are


- diagonal elements: Ykk is equal to the sum of all admittances with one end directly
connected to busbar k (including local loads represented by their shunt admittances rather
than by their MW and MVAr)
- off-diagonal elements: Ykn is equal to the negative value of the sum of all admittances
directly connected between busbars k and n ( k  n )

Using [Y ] (assumed known), we can write the nodal equation for a power network as

[ I ] = [Y ][V ]

- where [ I ] is the vector ( N  1 ) of busbar current phasors associated with local generators
(MW and MVAr) as well as local loads (MW and MVAr rather than if represented
by their shunt admittances)(excluding current flow through local loads represented
by shunt admittances avoids double counting)

[V ] is the vector ( N  1 ) of busbar voltage phasors (magnitude and angle)

54 ELEN5008A – Power Systems


For busbar k (now moving from a vector of current phasors and a vector of voltage phasors to
particular busbar current phasors and particular busbar voltage phasors), the total net current
phasor injection (LHS of equation) associated with local generators (MW and MVAr) as well as
loads (MW and MVAr rather than their shunt admittances) equals the current flow into local
shunt admittances as well as current flowing through series admittances from busbar k to the
other busbars (RHS of equation)( Ykn zero if there is no direct connection)( k = 1,2,...,N )
N
Ik =  YknVn
n =1

Note that one of the terms of the summation is for busbar k itself ( n = k ) which is YkkVk where
Ykk is the respective diagonal element in the network nodal admittance matrix (sum of all the
admittances directly connected to busbar k ).

For example, for a three-bus network

Y1 Y3
I 21 I 23

Bus 1 Bus 2 Bus 3


V1 V2 V3
I2 V2 Y22

Y2

I 2 = V2Y2 + I 21 + I 23

= V2Y2 + (V2 − V1 )Y1 + (V2 − V3 )Y3

= V1( −Y1 ) + V2 ( Y2 + Y1 + Y3 ) + V3 ( −Y3 )

= V1(Y21 ) + V2 (Y22 ) + V3 (Y23 )

3
Or I2 =  Y2nVn
n =1

Power flow / load flow analysis thus involves the simultaneous solution of the set of the N
nodal equations (from [ I ] = [Y ][V ] )
 S  N
Ik  = k = Y V
 V  n =1 kn n
 k 

ELEN5008A - Power Systems 55


- where k represents a node being considered from a total of N nodes
n represents all nodes when considering node k (including node k itself)
I k represents the net current phasor injected into node k by generators (MW and
MVAr) as well as loads (MW and MVAr rather than represented by their shunt
admittances)
Sk represents the net complex power injected into node k by generators (MW and
MVAr) as well as loads (MW and MVAr rather than represented by their shunt
admittances)

Note that if there is an admittance connected between node k and the reference node (node 0 )
(e.g. load connected to node k and represented as an admittance rather than as a complex power)
this appears in the Ykk admittance term.

The reference node (the neutral in a three-phase network) is excluded from the above N nodes.

j
For k = 1,2,...,N , n = 1,2,...,N Vk = Vk e k

j
Vn = Vn e n

Ykn = Gkn + jBkn

Then Sk = Pk + jQk = Vk I k
*
 N  N
= Vk   VnYkn  =  VkVnYkn

 n =1  n =1
N
j( k −n )
=  Vk Vn ( Gkn − jBkn )e
n =1
N
=  Vk Vn ( Gkn − jBkn )(cos( k − n ) + j sin( k − n ))
n =1

From the real and imaginary components, we can write the power flow equations as

N
Pk =  Vk Vn (Gkn cos( k − n ) + Bkn sin( k − n ))
n =1
N
Qk =  Vk Vn (Gkn sin( k − n ) − Bkncos( k − n ))
n =1

Contrast these equations with the equations in Chapter 1 which exclude n = k in the summation
(the equations in Chapter 1 are preferred when dealing with individual transmission lines).

56 ELEN5008A – Power Systems


Solution of the above set of equations is made more difficult because of the nonlinearity of the
equations – the powers, currents and voltages are all interdependent.

The method of solution must therefore be iterative – starting with an initial guess of the busbar
voltages (the bus voltage phasor magnitudes equal to their nominal values with phase angles all
zero is a good initial guess) and then making successive adjustments to the solution until the
solution converges at the final best-estimate of the bus voltage phasor magnitudes and phase
angles.

Two methods of iterative solution are used to solve the power balance equations
- Newton-Raphson method
- Gauss-Seidel method

One method of solution, involving inversion of the nodal admittance matrix, is considered next.

4.3.2 Direct method involving inversion of the nodal admittance matrix

The advantage of this method is that the nodal admittance matrix needs to be inverted only once
– later changes to the network (adding loads, changing component impedances) can be made
directly to the inverted matrix.

The direct method involves inverting the nodal admittance matrix [ Y ] so that the original nodal
matrix equation
[ I ] = [ Y ][V ]

- is replaced with [V ] = [Y ] −1 [ I ]

This requires that the nodal admittance matrix [ Y ] is square and non-singular (has an inverse) –
true for all power system nodal admittance matrices.

The iteration then involves


- the unknown nodal voltages are first estimated (initial guess is usually the nominal voltage)
- using these estimated nodal voltages and known complex powers at each node, a first set of
estimated currents injected into each node is obtained
- using these estimated injected currents and the inverted nodal admittance matrix, a corrected
set of nodal voltages is obtained
- using these new nodal voltages together with the known complex powers at each node, a
new set of estimated currents injected into each node is obtained
- this iterative process is terminated when successive values of calculated nodal voltages
converge such that the difference between two successive sets of nodal voltages lies within a
predefined limit

ELEN5008A - Power Systems 57


4.3.3 Modification of the inverse of the nodal admittance matrix [Y ] −1

This modification may be required due to the


- addition of a new load represented by an admittance (want to know whether the nodal
voltages are still acceptable and no network components are overloaded)
- removal of a transmission line (want to know if the nodal voltages are still acceptable and no
remaining network components are overloaded)
- addition of a transmission line
- changing tap in a transformer with a tap changer

This modification must not involve addition of new nodes or then [ Y ] needs to be
reformulated.

We define [ Y ] as the original nodal admittance matrix and [Ym ] as the modified nodal
admittance matrix (both n  n matrices).

Adding a new load admittance between node r and the reference node (node 0 ) then just
involves increasing the respective self-admittance term in [ Y ] .

The modified admittance matrix [Ym ] can be expressed as the sum of [ Y ] and a modification
matrix (both n  n matrices)

[Ym ] = [Y ] +  [ G ][ H ]t

-where  is a scalar
[ G ] is a column vector
[ H ]t is a row vector where [ H ] itself is a column vector

Then (since we eventually want the inverse of the modified admittance matrix when we already
have the inverse of the original admittance matrix)(proof can be found in the references)

 [Y ] −1 [ G ]   [ H ] [Y ] −1 
   
−1
=  [Y ] +  [ G ][ H ]t  −1 = [Y ] −1 − 
t
[Ym ]
+  [ H ]t [Y ] −1 [ G ] 
1
  

Two examples are given below


- addition of a new load (at a node)
- addition of a new transmission line (new impedance between two nodes)

58 ELEN5008A – Power Systems


4.3.3.1 Addition of a new load (new load admittance at a particular node)

Assume a new load admittance y is to be added to the network between node k and the
reference node (node 0 ).

The modification matrix for the calculation of [Ym ] will then have the following form (value of
kth diagonal term equal to y )

1 2 k n
1 0 0 . . . 0 . . . 0
2 0 0 . . . 0 . . . 0
. . . .
. . . .
 [ G ][ H ]t = . . . .
k 0 0 y 0
. . . .
. . . .
. . . .
n  0 0 . . . 0 . . . 0 

This implies that we can choose  = 1 and [ G ] = [ H ] so that

1 0 
2 0 
 . 
 . 
 . 
[G] = [ H ] =
k y
 . 
 . 

. 
n  0 

Thus
 [Y ] −1 [ G ]  gives column k of [Y ] −1 multiplied by y
 

 [ H ] [Y ] −1  gives row k of [Y ] −1 multiplied by y


 t 

 [ H ] [Y ] −1 [ G ]  gives element kk of [Y ] −1 multiplied by y


 t 

 [Y ] −1 [ G ]   [ H ] [Y ] −1 
   
[Ym ] −1 = [Y ] −1 − 
t
Thus
+ [ H ]t [Y ] −1 [ G ] 
1 
  

−1 [ column k of [Y ] −1 ][ row k of [Y ] −1 ]
= [Y ] −
1 −1
+ ( element kk of [Y ] )
y

ELEN5008A - Power Systems 59


4.3.3.2 Addition of a new impedance between two nodes

If a new impedance z is added to the network to interconnect nodes f and k , then the nodal
self-admittances at nodes f and k are increased by 1 / z and the nodal mutual admittances
between nodes f and k are reduced by 1 / z so that the modification matrix for the calculation
of [Ym ] is given by
f k
 . . 
 . . 
 . . 
 . . . 1 / z . . . −1 / z . . . 
f  . . 
 [ G ][ H ]t =  . . 
 . . 
 . . . −1 / z . . . 1 / z . . . 
k  . . 
 . . 
 
 . . 

This implies that we can choose  = 1 and [ G ] = [ H ] so that

 0 
 . 
f 1/ z 

 . 
[G] = [ H ] =  0 
 . 
k −1 / z 

 
 . 
 0 
Thus

 [Y ] −1 [ G ]  gives a column vector with each element consisting of an element of column f


 
minus the corresponding element of column k of [Y ] −1 divided by z

 [ H ] [Y ] −1  gives a row vector with each element consisting of an element of row f minus
 t 
the corresponding element of row k of [Y ] −1 divided by z

 [ H ] [Y ] −1 [ G ]  gives the sum of elements ( ff + kk − fk − kf ) of [Y ] −1 divided by z


 t 

 [Y ] −1 [ G ]   [ H ] [Y ] −1 
   
−1 −1
−
t
Thus [Ym ] = [Y ]
+  [ H ]t [Y ] −1 [ G ] 
1 
  

−1 −1
−1 [ column f − column k of [Y ] ][ row f − row k of [Y ] ]
= [Y ] −
−1
z + ( sum of elements ( ff + kk − fk − kf ) of [Y ] )

60 ELEN5008A – Power Systems


4.3.4. The Newton-Raphson iterative method

The Newton-Raphson iterative method is used to solve nonlinear equations.

For a nonlinear function of a single variable, f ( x ) , for iteration j , the Newton-Raphson


iterative equation for the estimate of the single unknown, x , x̂ , is

f ( x ) x = x( j −1 )
x( j ) = x( j − 1 ) − or x( j − 1 ) − x( j − 1 )
df ( x )
dx x = x( j −1 )

e.g. if f ( x ) = x3 + 2x − 2

df ( x )
= 3x2 + 2
dx

Assuming the first estimate is x(0 ) = 1

3
x + 2x − 2
x = x( 0 ) 1
Then x( 1 ) = x( 0 ) − = 1− = 0,8
2
3x + 2 5
x = x( 0 )

x( 2 ) = 0,7714

x( 3 ) = 0,7709

etc.

Thus in the Newton-Raphson iterative method

df ( x )
− f ( x ) x = x( j −1 ) =  −x( j − 1 )
dx x = x( j −1 )

df ( x )
-where − f ( x ) x = x( j −1 ) and are known and hence −x( j − 1 ) can be
dx x = x( j −1 )
calculated.

Then x( j ) = x( j − 1) + x( j − 1)

ELEN5008A - Power Systems 61


The Newton-Raphson iterative method can also be used to solve a set of n nonlinear equations
with n unknowns

f1( x1 ,x2 ,...,xn ) = b1


f 2 ( x1 ,x2 ,...,xn ) = b2

f n ( x1 ,x2 ,...,xn ) = bn

The iterations start with initial ( j = 0 ) estimates of the n unknowns

x1( 0 ),x2 ( 0 ),...,xn ( 0 )

- and corrections to the estimates with the aim of converging to the true values

x1( 0 ), x2 ( 0 ),..., xn ( 0 )

f1( x1( 0 ) + x1( 0 ),x2 ( 0 ) + x2 ( 0 ),...,xn ( 0 ) + xn ( 0 )) = b1


f 2 ( x1( 0 ) + x1( 0 ),x2 ( 0 ) + x2 ( 0 ),...,xn ( 0 ) + xn ( 0 )) = b2
Thus
f n ( x1( 0 ) + x1( 0 ),x2 ( 0 ) + x2 ( 0 ),...,xn ( 0 ) + xn ( 0 )) = bn

Each of the above equations is linearized using the Taylor series expansion (only the first term
for small deviations from the true value).

For equation 1 and for the first estimate ( j = 0 )

fi ( x1( 0 ) + x1( 0 ),x2 ( 0 ) + x2 ( 0 ),...,xn ( 0 ) + xn ( 0 ))

 f   f   f 
= fi ( x1( 0 ),x2 ( 0 ),...,xn ( 0 )) +  i  x1( 0 ) +  i  x2 ( 0 ) + ... +  i  xn ( 0 )
 x  x  x 
 1 x = x ( 0 )  2 x = x ( 0 )  n x = x ( 0 )
1 1 2 2 n n

= bi

In matrix form
 f1 f1 f1 
 
x x2 xn 
 b1 − f1( x1( 0 ),x2 ( 0 ),...,xn ( 0 ))   1  x1( 0 ) 
 b − f ( x ( 0 ),x ( 0 ),...,x ( 0 ))   f 2 f 2 f 2   x ( 0 ) 
 2 2 1 2 n  =  x x2 xn 
  2 
   1  
 bn − f n ( x1( 0 ),x2 ( 0 ),...,xn ( 0 ))     xn ( 0 ) 
 f f n f n 
 n 
 x1 x2 xn 
[ x = x( 0 )]

Or [ f ( 0 )] = [ J( 0 )][ x( 0 )]

62 ELEN5008A – Power Systems


[ J( j )] is known as the Jacobian matrix for iteration j (associated with linearization)(has to be
recalculated at each iteration).

For the particular iteration, all the elements in [ f ( j )] are known and all the elements in
[ J( j )] are known and hence all the elements in [ x( j )] can be calculated.

Updated values of the estimates [ x ] are then calculated from

[ x( j + 1)] = [ x( j )] + [ x( j )]

The iterations are stopped when all the elements in [ x( j )] drop to less than a predefined
tolerance.

When the Newton-Raphson iterative method is applied to the load-flow problem the active
powers Pk and reactive powers Qk are defined (superscript d ) for particular nodes (busbars)
and then the following equations are solved for the unknown node voltages and node voltage
phase angles

P1( 1( j ),..., n ( j ),V1( j ),...,Vn ( j )) = P1d

d
Pn ( 1( j ),..., n ( j ),V1( j ),...,Vn ( j )) = Pn

Q1( 1( j ),..., n ( j ),V1( j ),...,Vn ( j )) = Q1d

Qn ( 1( j ),..., n ( j ),V1( j ),...,Vn ( j )) = Qn d

 P1 P1 P1 P1 


 
 Pd ˆ ˆ   1  n V1 Vn 
− P1( 1( j ),..., n ( j ),V1( j ),...,Vn ( j ))
 1   
   Pn Pn Pn Pn   1( j ) 
Pd   
− Pn ( 1( j ),..., n ( j ),V1( j ),...,Vn ( j ))   1
ˆ ˆ  n V1 Vn   
 n

 
=    n ( j ) 
Q d   V ( j ) 
 1 − Q1( 1( j ),..., n ( j ),V1( j ),...,Vn ( j ))   Q1
ˆ ˆ Q1 Q1 Q1   1 
 
   1  n V1 Vn 
  Vn ( j ) 
 d 
 Qn − Qn ( ˆ1( j ),...,ˆn ( j ),V1( j ),...,Vn ( j ))   
 Qn Qn Qn Qn 
 
 1  n V1 Vn 
j

Or [ f ( j )] = [ J( j )][ x( j )]

ELEN5008A - Power Systems 63


Notes:

64 ELEN5008A – Power Systems


CHAPTER 5: POWER SYSTEM STABILITY

5.1 Introduction

Stability, when used with reference to a power system, is that attribute of the system which
enables it to develop restoring forces between its elements equal to or greater than any disturbing
forces so as to maintain a state of equilibrium between the elements.

Stability studies are considered in three parts


− steady-state stability (no disturbance present)
− small-signal stability (small disturbance present)
− transient stability (large disturbance present)

5.2 Steady-state stability

Steady-state stability refers to the behaviour of a power system where changes are made very
slowly.

The entire system can adapt itself to the changes and no overshoot of any parameter takes place.

Steady-state stability is only breeched if the power flow along a transmission line exceeds (as
shown previously)

VS VR
Plim =
Xs

where VS is the sending end rms line voltage, in V


VR is the receiving end rms line voltage, in V
X s is the series inductive reactance of the line, in Ω

5.3 Small-signal stability

Small disturbances excite rotor angle oscillations between generators.

The seriousness of these oscillations is determined by the amount of damping present in the
power system.

Small signal stability is not covered in this course.

ELEN5008A - Power Systems 65


5.4 Transient stability

Transient stability refers to the ability of a power system to regain equilibrium after a major
disturbance occurs in the power system (e.g. fault).

It is necessary to incorporate the rotational dynamics of the generators as well as the electrical
dynamics of the power system.

Consider a single generator interlinked to an infinite bus via a transmission line, as shown below

VS VR

Ts Te Xs

Generator

Infinite bus

Any change in electrical load or disturbance will manifest itself as a change in generator torques.

We define three generator torques

(a) Shaft torque, Ts - supplied by the prime mover (e.g. steam turbine, water turbine) and
will be assumed to remain constant for the duration of the disturbance

(b) Electromagnetic torque, Te - this is the retarding torque associated with the electrical
power output of the generator.

For the case where the power out of the generator flows along a transmission line

VS VR
Pe = sin 
Xs

- where  is the phase angle between the voltage phasors VS and VR .

(c) Acceleration torque, Ta - this is simply the difference between the shaft torque and the
electromagnetic torque
Ta = Ts − Te

If the shaft torque is greater than the electromagnetic torque, the generator will accelerate and if
the electromagnetic torque is greater than the shaft torque, the generator will decelerate.

66 ELEN5008A – Power Systems


Mechanical power can be obtained by multiplying the mechanical torque by the mechanical
angular velocity,  , which is assumed to vary only slightly during the disturbance - due to the
large shaft inertia.

V V
Thus Pa = Ps − Pe = Ps − S R sin 
Xs

where Pa is the acceleration power, in MW


Ps is the shaft power (mechanical power input), in MW
Pe is the electromagnetic power (electrical power output), in MW

VS , VR , X S and  have their usual meaning (voltages in kVline )

The above equation assumes mechanical losses and electrical losses in the generator are small.

2 2
d d  d 
Also (mechanical equations) Pa = Ta = J  = J 2 = M 2
dt dt dt

where  is the angular velocity of the generator shaft, in rad/s


J is the moment of inertia of the generator and prime mover shafts, in kg-m2
 is the mechanical angle of the rotor, in rads
M is the angular momentum of the generator and prime mover shafts ( = J  )

2
d 
Substituting into the previous equation M = Ps − Pe
dt 2

2
d 
Also written M = Ps − Pe
dt 2

- where  represents the disturbance in  .

This is known as the swing equation (describes how the angle of rotation (normally increases
linearly with time as the rotor rotates at constant speed) gets disturbed by the difference between
the mechanical power in and the electrical power out (implies speed no longer constant during
the disturbance)).

This is a second order differential equation with non-constant coefficients and cannot be solved
in closed form. Numerical techniques used to solve this equation are described later in the
chapter.

However, this equation can be solved graphically which also leads on to the equal area criterion
for return to equilibrium.

Question for students: what if we replace one generator connected to an infinite bus (through the
above inductive reactance) with two generators connected through the same inductive reactance?

ELEN5008A - Power Systems 67


5.4.1 Equal area criterion

d
Multiplying both sides of the swing equation by we get
dt

2
d d  d
M = ( Ps − Pe )
dt dt 2 dt

d  d 
2
d  d 2 
But (differentiation by parts)   =2  
dt  dt  dt  dt 2 
 

2
1 d  d  d
Thus M   = ( Ps − Pe )
2 dt  dt  dt

2
M  d 
Or   =
2  dt   ( P − P ) d
s e

d
 ( P − P ) d
2
Or = s e
dt M

d
For a turning point in  (representing a return to equilibrium) we need =0 .
dt

Or

disturbance in 
( Ps − Pe ) d = 0

- at the turning point, the areas where Ps  Pe must equal the areas where Ps  Pe - equal area
criterion.

For the case of one generator swinging relative to an infinite busbar through a transmission line
of inductive reactance X s
 

V V
 P − S R sin   d = 0
 s Xs 
 

- where  is the angle between VS (sending end transmission line voltage phasor) and VR
(receiving end transmission line voltage phasor).

If the two powers is plotted versus  , the above equation has the form of the area between the
Ps and the Pe curves (area either positive ( Ps  Pe ) or negative ( Ps  Pe )).

68 ELEN5008A – Power Systems


Returning to equilibrium after a disturbance requires that the positive area ( Ps  Pe ) must be
larger than the negative area ( Ps  Pe ).

Example

Consider a generator connected to an infinite busbar through two identical parallel lines, each
having a series inductive reactance X s

VS VR
X1 = Xs

Ps Pe

X2 = Xs
Generator

Infinite bus

Plotting Ps (assumed constant during the disturbance) and Pe as a function of 

P
VS VR
Pe1 = 2 sin  (both lines in service)
Xs
VS VR
Pe2 = sin  (only one line in service)
Xs

Ps


1 2 3

With both lines in service, the operating point will be at  1 and with only one line in service the
operating point will be at  2 .

ELEN5008A - Power Systems 69


Assume both lines are in service and then one line is disconnected – this implies  must
increase from  1 to  2 with overshoot to  c and then return, eventually after a few oscillations,
to  2 .

A1 = A2

Provided  c   3 the generator will eventually return to steady-state operation at  2 .

Another scenario is when there is a balanced three-phase short circuit on one line ( Pe = 0 for a
three-phase short circuit) – there must be a load with resistance for real power (megawatts) to
flow out of the generator.

Because Ps  Pe ( = 0 ) during the disturbance, the generator shaft accelerates (  increases).

We assume the three-phase short-circuit is cleared (at  4 ) by tripping the line with the three-
phase short-circuit and then operating just on the remaining line

70 ELEN5008A – Power Systems


A1 = A2

Provided  c   3 the generator will eventually return to normal operation at  2 .

This also implies the existence of an A2(max) .

For the first example of two parallel lines with no fault

Pe1

Pe2
A C A 2 (max) D
Ps
A1


1 2 3

Thus if A2(max)  A1 the system will be stable (will return to new equilibrium condition
after the disturbance)
A2(max)  A1 the system will be unstable (  continues to increase)

ELEN5008A - Power Systems 71


Example: Consider a double-circuit 400 kV transmission line having a series inductive reactance
of 30  per circuit (this would correspond to a line approximately 100 km in length).

The steady-state power limit (  =  / 2 ) per circuit would then be

400000 2
Pe(max)1 = = 5,33 GW
30

400000 2
For both circuits together Pe(max)2 = 2  = 10,66 GW
30

A steady-state situation initially exists with a load of 4,5 GW being fed via both circuits.

This requires that Ps = 4,5 GW

If one circuit is disconnected (the scenario being considered), the full load of 4,5 GW is
transferred to the remaining circuit.

Thus the equal area criterion states that the threshold of instability will be if

A2( max ) = A1

3
But A2(max) =  ( 5,33 sin  ) d − 4,5(  3 −  2 ) in GW-rad
2
= − 5,33 (cos  3 − cos  2 ) − 4,5 (  3 −  2 ) in GW-rad

A1 = 4,5 (  2 − 1 ) − ( −5,33 (cos  2 − cos 1 )) in GW-rad

But (from the waveforms)

 4,5 
1 = sin −1   = 24,97 or 0,436 radians
 10,66 

 4,5 
 2 = sin −1   = 57,59 or 1,01 radians
 5,33 

 3 = 180 −  2 = 122,41 or 2,14 radians

72 ELEN5008A – Power Systems


Thus A2(max) = ( 5,33 GW  1,072 − 4,5 GW  1,13 ) = 0.629 in GW-rad

A1 = ( 4,5 GW  0,574 − 5,33 GW  0,37 ) = 0,611 in GW-rad

Thus A2( max )  A1

Thus, this scenario results in  2 being eventually reached in a stable manner.

ELEN5008A - Power Systems 73


5.4.2 Numerical solution of the swing equation

The full swing equation, including a damping term, is

2
d  d
M = Ps − Pe(max) sin  − f
dt 2 dt

- where f is the damping coefficient, in MW-s/rad.

Normalizing the swing equation with respect to the angular momentum, M , we get

2
d  P Pe(max) f d
= s − sin  −
2 M M M dt
dt

d
= P's − P'e(max) sin  − f '
dt

Expressing in terms of two first order equations (replacing  with y to simplify the code)

y1 = 

y2 = y1

y2 = P's − P'e(max) sin y1 − f ' y2

74 ELEN5008A – Power Systems


CHAPTER 6: HVDC TRANSMISSION

6.1 The three-phase fully controlled bridge

We assume that the DC load is sufficiently inductive to maintain a constant DC current iDC and
we assume the AC supply impedance is zero.

We define the delay angle  as the angle between when the incoming SCR becomes forward
biased and when the incoming SCR is actually triggered ( 0     ) .

The respective waveforms are shown on the next page, where

(a) are the two DC terminal (+ve and –ve) voltages referenced to the AC supply neutral point
(b) is the voltage between the two DC terminals
(c) are the upper SCR currents
(d) are the lower SCR currents
(e) is one of the AC supply currents (red phase)

ELEN5008A - Power Systems 75


Waveforms – rectification mode ( 0     / 2 )

76 ELEN5008A – Power Systems


Waveforms – inversion mode (  / 2     )

ELEN5008A - Power Systems 77


If the AC line voltage is 2 VL cos t then the average DC voltage is calculated from (origin at
peak of line voltage waveform)


+
6 6
2  
VDC = 2 VL cos t d t
− +
6

- where VL is the RMS AC line voltage, in V.

3 2    
This gives VDC = VL  sin( +  ) − sin( − +  ) 
  6 6 

3 2      
= VL  sin cos  + cos sin  − sin( − )cos  − cos( − )sin  
  6 6 6 6 

3 2
= VL cos 

- VDC is positive for 0     / 2 (rectification mode) and negative for  / 2     (inversion


mode).

The average DC voltage as a function of  is as follows

78 ELEN5008A – Power Systems


6.2 HVDC transmission schemes

With HVDC (High Voltage DC) transmission, power is transferred via a DC link with a fully
controlled bridge operating in the rectification mode at the one end (power sending end) and a
fully controlled bridge operating in the inversion mode at the other end (power receiving end).

Advantages
- no stability problems
- fewer conductors (two versus three) and hence smaller pylons
- no skin effect - hence smaller conductors can be used
- no capacitive current
- asynchronous link (systems operating with different frequencies can be linked)
- short circuits can be detected and cleared rapidly

Disadvantages
- cost of converter equipment at both ends
- difficulty of switching DC
- difficulty of intermediate tap-off points

In practice the three-phase fully controlled bridges (6-pulse connection) are replaced with two
series-connected three-phase fully controlled bridges (12-pulse connection).

Because the two series-connected three-phase fully controlled bridges are supplied by two three-
phase voltage sets displaced by 30º (one fed off a star-connected secondary winding and the
other fed off a delta-connected secondary winding) this causes useful cancellation of the lower
order harmonics on the AC side (cancel 5th and 7th, 17th and 19th , etc. – leaving the 11th and 13th,
23rd and 25th, etc) and cancellation of particular harmonics on the DC side (cancel 6 th, 18th, etc. –
leaving the 12th, 24th, etc).

ELEN5008A - Power Systems 79


The 5th harmonic component is a negative-sequence harmonic (phasor rotates in the opposite
direction to the fundamental frequency phasor) and hence when referenced to the fundamental
frequency phasor it rotates at six times the angular velocity of the fundamental frequency phasor.
Similarly the 7th harmonic component is a positive-sequence harmonic (phasor rotates in same
direction as the fundamental frequency phasor) and hence when referenced to the fundamental
frequency phasor it rotates at six times the angular frequency of the fundamental frequency
phasor.

When the fundamental frequency phasor is shifted by 30° both the 5th and 7th harmonics are
shifted by 6×30° = 180°. When the shifted waveform is added to the unshifted waveform the
resultant waveform does not have 5th and 7th harmonics (sum to zero).

Series connection of two three-phase bridges to give a 12-pulse connection

Thus the full HVDC scheme is more likely to be as shown below

80 ELEN5008A – Power Systems


CHAPTER 7: BASIC PRINCIPLES OF PROTECTION

7.1 Introduction

Faults (short circuits) in power systems are inevitable and hence equipment must be installed to
− detect that a fault has occurred (e.g. by monitoring the currents)
− initiate the opening of a circuit breaker as quickly as possible to limit, due to the large fault
currents
− damage at the fault location
− mechanical damage to upstream transformers due to large electromechanical forces
when the short-circuit current flows
− thermal damage to upstream insulation such as in cables due to the large heat generation
when the short-circuit current flows

The fault clearance must also affect as few customers as possible - only the faulted circuit must
be isolated.

Overload (rather than fault) conditions can also be threatening to equipment (transformers,
motors, etc) and hence must also be limited in duration.

Circuit breakers used to interrupt the fault current must be suitable rated
− speed of opening following trip signal from protection relay (typically > 60 ms)
− ability of contacts to interrupt the large fault currents that can flow in power systems

Rupturing capacity of circuit breaker = 3  nominal line voltage  fault current rating

Network fault level = 3  nominal line voltage  actual fault current

The table below gives typical ratings for circuit breakers used in power systems (rated current
typically < few kA)

Nominal line voltage Network fault level Fault current rating


11 kV 150 MVA – 500 MVA 16 kA – 50 kA
66 kV 1500 MVA – 2500 MVA 16 kA – 50 kA
132 kV 2500 MVA – 5000 MVA 16 kA – 50 kA
275 kV 10000 MVA – 15000 MVA 25 kA – 50 kA
400 kV 17500 MVA – 35000 MVA 25 kA – 50 kA

Obviously we must have

Rupturing capacity of circuit breaker > network fault level at circuit breaker

Fault current rating of circuit breaker > fault current at circuit breaker

ELEN5008A - Power Systems 81


Typical protection relays include
− instantaneous overcurrent relays – if the fault current exceeds some threshold value the relay
immediately initiates the tripping of a circuit breaker
− definite-time overcurrent relays – as above but with an additional constant delay
− inverse-time overcurrent relays – the higher the overload current the quicker the circuit
breaker is tripped – to match the typical overload and fault current capability of power
equipment
− directional overcurrent relays – by sensing both voltage and current the direction of the fault
can be determined – useful where the power network has the form of a mesh (fault current
can flow in either direction round the mesh)
− directional inverse-time overcurrent relays – combination of the previous two
− undervoltage relays – if the supply voltage is too low this can be hazardous to motors and
hence a trip is initiated after a constant delay
− differential current relays – if the current flowing out of a circuit (e.g. transformer) differs
from what is expected from the current flowing into the circuit this suggests an internal fault
and a trip is initiated
− impedance relays – by sensing both voltage and fault current, the location of the fault can be
calculated (from the known circuit impedances) and if it is inside the zone-of-protection of
the relay the circuit breaker is tripped

The design of protection schemes involves the following


− determine which types of relays are needed
− supply the relays with the required currents and voltages (suitably reduced by current
transformers (CTs) and voltage transformers (VTs) to what can be managed by protection
relays)
− determine the relay settings to give the required tripping characteristic

Two types of (low VA rating) instrument transformers are used to monitor power systems
− metering transformers (e.g. for measuring energy consumption) – must be very accurate
− protection transformers (for detecting fault conditions) – must be reasonably accurate
particularly under fault conditions

82 ELEN5008A – Power Systems


7.2 Conventional Voltage Transformers (VTs)

The standard secondary line voltage for VTs is 110 V (RMS); therefore typical line voltage
ratios of VTs are 11000/110 V, 22000/110 V, 275000/110 V, etc.

At the lower voltage levels ( 132 kV) the VT is the conventional double-winding
electromagnetic transformer and is connected either phase-phase or phase-earth.

A typical connection together with the phasor diagram is shown below

To sense earth faults, the residual voltage is also measured

If the individual phase-neutral voltages are measured, the residual voltage can also be calculated
(modern microprocessor-based protection relays).

At the higher voltage levels (>132kV) the cost of a conventional VT with an HV-insulated
winding would be prohibitive and hence a lower voltage VT is used together with a capacitor
divider (known as a capacitor-coupled VT or CVT) - always connected phase-earth

ELEN5008A - Power Systems 83


CVTs can only measure 50 Hz voltages accurately (cannot easily be used to measure harmonic
voltages).

The capacitor CHV can also be used to inject high frequency carrier signals onto the power line
(for Power Line Carrier (PLC) based protection relay schemes – see later).

Characteristics important for protection VTs include


− accurate voltage ratio even up to 1,9 pu power frequency overvoltage conditions that could
be present during the fault (designed with lower flux densities to avoid saturation at the
higher power frequency overvoltage levels)
− low phase angle error (important when measuring impedance in conjunction with current
transformers)

84 ELEN5008A – Power Systems


Metering class VTs (0,8 pf lagging burden)

Accuracy class Maximum voltage magnitude error Maximum voltage phase angle error
0,1  0,1%  5 minutes
0,2  0,2%  10 minutes
0,5  0,5%  20 minutes
1,0  1,0%  40 minutes
3,0  3,0% Not specified

Protection class VTs (P for protection)

Accuracy class Maximum voltage magnitude error Maximum voltage phase angle error
3P  3%  120 minutes (2)
6P  6%  240 minutes (4)

VTs are also rated in terms of rated apparent power output (e.g. 50 VA) at rated output voltage
(e.g. 110 V).

7.3 Conventional Current Transformers (CTs)

There are two standard secondary currents for CTs: 1 A (RMS) and 5 A (RMS); therefore typical
current ratios are 100/1 A, 100/5 A, 10000/1 A, etc.

There are two CT types


− bar primary – single turn primary (busbar inserted through center of toroidal core)
− wound primary – multiple turn primary (smaller primary currents)

CT secondary windings can be connected in a star configuration (measure the line currents as
well as the sum of the three line currents (residual current – zero except when there is an earth
fault))

The protection relay input impedances (convert the CT secondary current to a (very small)
voltage) are represented by the impedances (rectangles) above.

ELEN5008A - Power Systems 85


The CT secondary windings can also be connected in a delta configuration

With the delta configuration, the residual current flows round the loop consisting of the CT
secondary windings and does not flow through the relay input impedances (the relay will not be
able to detect earth faults).

CTs need to measure currents accurately over a very wide range – up to the expected fault
current – which can be many times larger than the rated current of the equipment being
protected.

There are two issues concerning CT behaviour


− the CT has very little effect on the value of current flowing in the primary circuit and hence
the primary current can be considered to be a current source, with the result that the primary
circuit impedances (primary winding resistance and primary winding leakage inductance)
can be ignored in the equivalent circuit – do not affect the value of secondary current
− the CT is deliberately designed to have a very small secondary leakage inductance, L2

Thus the equivalent circuit of a CT is as follows (as seen from the secondary side)

The current source represents the primary current referred to the secondary side.

The magnetizing current and core loss current introduce an error between the actual secondary
current and the secondary current predicted from the CT turns ratio.

86 ELEN5008A – Power Systems


The magnitude of the magnetizing and core loss currents is dependent on the secondary voltage
of the CT and hence on the magnitude of the burden impedance (should be as small as possible –
ideally a short-circuit).

A protection class CT is specified in terms of its


− rated output (in VA) at rated output current (1 A or 5 A)
− accuracy class (P for protection)
− upper current limit beyond which accuracy is not guaranteed (accuracy limit current)

Accuracy Maximum current magnitude Maximum composite Maximum current phase angle
class error at rated primary current error at accuracy error at rated primary current
limit current
5P 1% 5%  60 minutes (1)
10P 3% 10% Not specified

Define the composite error,  c ( K n is the CT turns ratio) – rms value of the instantaneous
current error values calculated over one power frequency cycle – expressed as a percentage of
the rms value of the primary current

T
100 1

2
c = ( K nis − i p ) dt
I p( rms ) T
0

Thus for a 1000/1 protection CT with a 15 VA-10P20 designation


− rated primary current is 1000 A
− rated secondary current is 1 A
− rated output is 15 VA (15 V at 1 A)
− rated burden impedance is 15  (15 V  1 A)
− accuracy class is 10P
− accuracy limit current factor is 20 (accuracy limit current of 20 kA)
− rated secondary voltage at accuracy limit current is 20  15 V = 300 V
− magnetizing current at accuracy limit current will not exceed 10% of 20 A or 2 A

Typically the above CT will never be operated (under worst case fault conditions) such that the
secondary voltage exceeds 0,25  300 V – provides some margin for DC offsets.

Class X protection CTs are used for high impedance differential protection (discussed later)
where the CT behaviour (error due to magnetizing current component) must be accurately
known.

ELEN5008A - Power Systems 87


The Class X protection CT is always of the bar type and thus has a very small secondary leakage
inductance.

Class X protection CTs are specified in terms of


− maximum secondary winding resistance, R2
− magnetization current I m as a function of the secondary voltage (magnetization curve)

Consider the following example


− 1000/1 CT ratio
− relay setting of 0,2 A (200 A primary current if CT were ideal)
− CT secondary resistance (including resistance of connecting leads) of 1 
− relay burden resistance of 75 
− CT secondary leakage inductance can be ignored
− core loss current I c can be ignored

Thus when the relay is set to trip at a CT secondary current of 0,2 A

Total secondary resistance = (75+1)  = 76 


Voltage across magnetization impedance at 0,2 A = (76  0,2) = 15,2 V
Magnetizing current (from magnetization curve) = 90 mA
Total equivalent CT secondary current = 0, 22 + 0, 092 = 0,219 A
Effective primary current setting = 0,219  1000 = 219 A (not 200 A)

Modern protection relays have very low burden resistances to minimize the risk of CT saturation.

Saturation is an important consideration when iron-cored CTs are used.

Saturation of CTs usually occurs for situations where the fault current has a decaying DC
component (combination of resistance and inductance upstream of the fault – always the case)

88 ELEN5008A – Power Systems


The DC component of the CT secondary voltage causes the flux to move into the saturation part
of its magnetization curve

With a reactor, too much current saturates the core; however with a transformer there are two
windings and the flux produced by the secondary winding current opposes the flux produced by
the primary winding current – we need another way of determining whether the core is saturated
– use the secondary winding voltage, v .

1 1
Flux excursion =
N sec  v dt = N ( volt-second )
sec

Peak flux excursion


Peak flux density, B =
Core area

The decaying DC component in the primary current gets reflected as a decaying DC component
in the voltage across the secondary burden resistance.

ELEN5008A - Power Systems 89


The secondary voltage waveform can therefore be resolved into two components
− sinusoidal AC component
− decaying DC component

The secondary volt-second can therefore also be resolved into two components

− due to the AC component (integration over a quarter cycle gives the peak flux excursion)

T /4  /2
1 1  /2 1
 sin t dt =
  sin t dt = −

cos t|
0
=

0 0

− due to the decaying DC component (integration to infinity gives the peak flux excursion)

 − Rt
L

e L dt =
R
0

Thus the total volt-second at the peak flux excursion is

L X
+1 +1
L 1
+ = R = R
R   

X
Thus the ratio has the form of a per-unit (peak AC component represents 1pu) increase in the
R
peak flux excursion.

X
Since a typical value of the ratio is 12, the peak flux excursion is 13 times that for when only
R
the AC component is present.

This emphasizes the threat that the DC component (always present in AC fault current
waveforms) poses to CTs.

Other CT issues are


− the secondary winding of a current transformer must never be open circuited (with current
flowing in the primary circuit) since the open circuit secondary voltage can reach a value
dangerous to the CT insulation (remember that the CT has many more secondary turns than
primary turns – step down current but step up voltage)
− there is also a problem when the circuit breaker opens – all the CTs remain magnetized
(remnant magnetism) and the flux locus is offset when the CTs are re-energized (remnant
magnetism is eventually self-eradicated)

90 ELEN5008A – Power Systems


7.4 Optical current transformers (OCTs)(non-conventional CTs)

Optical current transformers use the Faraday effect – a magnetic field causes particular types
of crystals to become birefringent (slightly different refractive indices for left-hand and right-
hand circularly polarized light).

Linearly polarized light may be considered as a vector sum of circularly polarized light in
both directions of rotation with equal amplitudes
- left-hand circularly polarized
- right-hand circularly polarized (the direction of rotation of the vector is related by the
right-hand rule to the direction of propagation)

Vertically polarized light – start of path through birefringent material (direction out of the
page):

Vector
sum

right left

ELEN5008A - Power Systems 91


Polarized light – end of path through birefringent material (direction out of the page):

Vector
sum

right left

As shown the right-hand circularly polarized


light is assumed to have a higher speed
(smaller refractive index)
Speed = Wavelength × Frequency

The Faraday effect thus produces rotation of a linearly polarized light beam:

92 ELEN5008A – Power Systems


The Faraday effect can be expressed as  = Ve B

- where Ve is the Vernet constant of proportionality


B is the axial flux density, in T
is the axial length, in m

The magnetic field at radius r encircling a conductor carrying current I is

 I
B= o
2 r

If the optic fiber encircles the conductor N times

 I  I
 = Ve o = Ve o 2 rN = Ve o I  N
2 r 2 r


Or I=
Ve o N

The measurement set-up is as follows:

Light source, Iin


Polarizer
(90º)

Faraday sensor

Analyzer
(45º)

Detector

Output, I out

ELEN5008A - Power Systems 93


The detector output intensity is (Malus’ law)

Iout = k AIin cos 2 ( 45 −  )

1 + cos 2( 45 −  ) 1 + cos( 90 − 2 ) 1 + sin( 2 )


But cos 2 ( 45 −  ) = = =
2 2 2

1 + sin( 2 )
Thus Iout = k A Iin
2

 2I out − k AIin 
 = sin−1 
1
Thus 
2  k A Iin 
 

Combining with the previous equation

1 −1  2I out − k A Iin 
I= sin  
2Ve o N  k A Iin 
 

7.5 Optical voltage transformers (OVTs)(non-conventional VTs)

Optical voltage transformers use the Pockels effect – an electric field causes particular types
of crystals to become birefringent (slightly different refractive indices for linearly polarized
light aligned with the Pockels sensor crystal and perpendicular to the Pockels sensor crystal).

Linearly polarized light


E in
E in(X) E in(Y)

Y
X
Pockels sensor
E out(Y)
E out(X)

 V

94 ELEN5008A – Power Systems


The measurement set-up is as follows:

Voltage
V
Light source, I in
Polarizer
X Y (90º)
Quarter-wave Pockels sensor
retardation plate XY
d
Analyzer
(0º) 

 + 90º
Detector

Output, I out

 V 
Define  =  
V 
  

- where V is the half-wave voltage (the voltage for which the phase shift is half a
wavelength, or  )

d
Define V =
3
2r220

- where  is the wavelength of the light, in m


d is the width of the Pockels sensor, in m
r22 is the electro-optic coefficient (Pockels coefficient) for the respective crystal
0 is the refractive index of the crystal with no electric field
is the length of the Pockels sensor, in m

At the output of the analyzer (Malus’s law applied to the X-component and the Y-component
of the linearly polarized light)

Iout = k AIin (cos 2 45 + cos 2(135 −  − 90 ))

 1 + cos 2( 45 ) 1 + cos 2( 45 −  ) 


= k A Iin  + 
 2 2 

ELEN5008A - Power Systems 95


 1 + cos 90 1 + cos ( 90 − 2 )
= k A Iin  + 
 2 2 

 sin ( 2 ) 
= k A Iin  1 + 
 2 

 2I out − 2k A Iin 
 = sin−1 
1
Thus 
2  k A Iin 
 

Combining with the previous equation

V −1  2I − 2k AIin 
V =  sin  out 
2  k A Iin 
 

Because V is of the order of a few kV, for measuring higher voltages the Pockels sensor is
connected to the output of a capacitor voltage divider.

96 ELEN5008A – Power Systems


CHAPTER 8: PERCENT AND PER UNIT QUANTITIES

8.1 Introduction

When performing power system calculations the main interests are usually
− percent (or per unit) volt drops during normal operating conditions
− percent (or per unit) overcurrents during faults

Also percent (or per unit) voltages, currents and impedances are referred unchanged across
transformers (a huge advantage in analyzing power networks containing transformers (multiple
voltage levels) – from UHV to HV to MV to LV).

The use of a percent (or per unit) voltage or a percent (or per unit) current implies the existence
of a base voltage and a base current (representing 100% or 1 pu voltage or current).

First, a base MVA is chosen (the base MVA is a true base value – it is the same value for all
voltage levels throughout the network)(e.g. 100 MVA or some other value that could have
particular significance in the particular power network – e.g. a supply transformer MVA)

VAbase = 3 Vbase Ibase

Second, the base voltage is chosen to be the nominal line voltage of the section of the network
being considered (e.g. 400 V, 11 kV, 400 kV).

Third, the base current is calculated from the base voltage using the above equation (varies with
the base voltage of the section of the network being considered).

Note that while Vbase is the rated nominal voltage in a part of the network, Ibase is not
necessarily related to any rated current of any component in the network at that nominal voltage.

Finally, the base impedance of the section of network being considered is calculated from

Vbase Vbase2 ( kVbase )2


Zbase = = or
3 Ibase VAbase MVAbase

Thus the base quantities (of the section of network being considered) are interrelated as follows
(for a three-phase network)
V base
3 Z base I base

VA base = 3 V base I base


V base
V base2
=
Z base

= 3 I base2 Z base

ELEN5008A - Power Systems 97


Z
For any impedance Z% = actual  100%
Zbase

Z
Similarly Z pu = actual
Zbase

The presence of impedance (in particular capacitive and inductive reactance) implies that there is
a base (angular) frequency, base (typically 50 Hz or 100π radians per second).

Similar equations are used to determine the percent or per unit voltages and currents from the
actual voltages and currents (assume have already determined the respective Vbase and Ibase )

If we know the fault level at a particular bus in the network (the usual situation) we can derive
the Thévenin equivalent of the upstream network as seen at this bus

Z %(supply)

V = 100%

Bus

MVAbase
- where Z%( supply ) =  100%
MVA fault −level

For example if the utility states that the fault level at the voltage level being considered is
MVA fault −level = 1000 MVA and we are assuming MVAbase = 100 MVA then the percent supply
impedance is 10% in our calculations.

To convert a base impedance (in Ω) at one voltage level ( kV previous ) to a base impedance (in Ω)
at another voltage level ( kVnew )
2
 kV 
Zbase( kV = Z  new 
new ) base( kV previous )  kV 
 previous 

If kVnew  kV previous then Zbase( kV  Zbase( kV - just as with transferring actual


new ) previous )
impedances (in ) from the LV side to the HV side of a transformer.

An impedance expressed as a percent (or per unit) impedance remains unchanged when moving
across a transformer because both the actual impedance (numerator) and the base impedance
(denominator) are multiplied by the same square of the transformer turns ratio which cancel.

To convert a percentage impedance expressed in one MVAbase ( MVAbase(previous) ) to the


percentage impedance in another MVAbase ( MVAbase( new ) )

98 ELEN5008A – Power Systems


 MVAbase( new ) 
 
Z%( MVA
base( new ) ) = Z%( MVA base(previous) )  MVA 
 base(previous) 

 MVAbase( new ) 
 
Similarly Z pu( MVA
new ) = Z pu( MVA previous )  MVA 
 base(previous) 

This is usually necessary with transformers since the transformer percent impedance (nameplate
data) is specified using the transformer MVA rating as the base MVA (which is the usual case
when the transformer is considered in isolation).

Thus we must make the conversion


 MVAbase 
 
Z%( MVA
base ) = Z%( MVA rated( transformer ) )  MVA 
 rated( transformer ) 

For example for a 1 MVA transformer with a nameplate percent impedance of 8% and where the
base MVA is 100 MVA
 100 
Z%( 100MVA ) = 8%   = 800%
 1 

With a 1 MVA transformer with an impedance of 8% this corresponds to having to apply a


voltage of 8% rated voltage to the primary winding with the secondary winding short circuited
to get the transformer rated current to flow through the transformer.

With a 1 MVA transformer with an impedance of 800% (now converted to the base MVA of
100 MVA) this corresponds to having to apply a voltage of 800% rated voltage to the primary
winding with the secondary winding short circuited to get the (typically much higher) base
current (no longer the transformer rated current) to flow through the transformer.

Once all percent impedances are related to the same MVA base they can be simply added (the
advantage of using percent impedances – can forget about the voltage change by the
transformers)(assume all impedances are inductive reactances so we do not need to use
Pythagoras’s theorem)

10% 50% 20% 800% 100%


V = 100%

Bus A

980%
V = 100%

Bus A

Then the three-phase short circuit current at Bus A will be

100%
 100% or 10,2% or 0,102 pu or 0,102  Ibase( Bus A ) in A
980%

ELEN5008A - Power Systems 99


Notes:

100 ELEN5008A – Power Systems


CHAPTER 9: SYMMETRICAL COMPONENTS

9.1 Introduction

Currents flowing in the different phases during unbalanced faults can be more easily calculated
by transforming to balanced three-phase currents (symmetrical components).

The symmetrical components approach is valid only if the network impedances are balanced
(only then does one sequence current have no effect on the others).

Unbalanced three-phase voltages or currents can be resolved into the sum of three sets of
balanced voltages or currents
- positive sequence (subscript 1) – three equal amplitude sinusoidal waveforms displaced by
120 and with the same phase rotation as if the original voltages and currents were balanced
- negative sequence (subscript 2) – three equal amplitude sinusoidal waveforms displaced by
120 and with the opposite phase rotation (red unchanged but white phase and blue phase
swapped)
- zero sequence (subscript 0) – three equal amplitude sinusoidal waveforms all in phase

This transformation is shown diagrammatically below (choosing the red phase as the reference
phase)

Having the phasors rotate in the same direction at the same angular velocity allows single phase
analysis of combinations of the positive, negative and zero sequence components.

Thus adding the components (phasor sum)

I R = I R1 + I R2 + I R0

IW = IW 1 + IW 2 + IW 0

I B = I B1 + I B2 + I B0

ELEN5008A - Power Systems 101


But from the sequence phasor relationships

( a = 1120, a2 = 1240 and a3 = 1360 = 1 )

I B1 = aI R1

IW1 = a2 I R1

IW 2 = aI R2

I B2 = a 2 I R2

I R0 = IW 0 = I B0

Therefore I R = I R1 + I R2 + I R0

IW = IW1 + IW 2 + IW0 = a2 I R1 + aI R2 + I R0

I B = I B1 + I B2 + I B0 = aI R1 + a2 I R2 + I R0

 IR   1 1 1  I 
Or I  =  a2 a 1   I R2 
R1
 W   2  
 IB   a a 1   I R0 

Rewriting the above I R = I R1 + I R2 + I R0

aIW = aIW1 + aIW 2 + aIW0 = I R1 + a2 I R2 + aI R0

a2 I B = a2 I B1 + a2 I B2 + a2 I B0 = I R1 + aI R2 + a 2 I R0

Also since I R2 + a 2 I R2 + aI R2 = 0

I R0 + aI R0 + a2 I R0 = 0

Thus 3I R1 = I R + aIW + a 2 I B

102 ELEN5008A – Power Systems


Similarly a2 I R = a2 I R1 + a2 I R2 + a 2 I R0

aIW = aIW1 + aIW 2 + aIW0 = I R1 + a2 I R2 + aI R0

I B = I B1 + I B2 + I B0 = aI R1 + a2 I R2 + I R0

Thus 3a2 I R2 = a2 I R + aIW + I B

Or 3I R2 = I R + a2 IW + aI B

Similarly I R = I R1 + I R2 + I R0

IW = IW1 + IW 2 + IW0 = a2 I R1 + aI R2 + I R0

I B = I B1 + I B2 + I B0 = aI R1 + a2 I R2 + I R0

Thus 3I R0 = I R + IW + I B = I N

 I R1  1 a a2   I 
I  = 1  R
a   IW 
2
To give
 R2  1 a
3 1 1 1  I 
 I R0    B 

The neutral current, I N , is three times I R0 since the zero sequence currents of all three phases
flow through the neutral and the three zero sequence currents have equal amplitude and are in
phase.

ELEN5008A - Power Systems 103


9.2 Sequence network relationships during faults

Just as the positive sequence currents see particular network impedances, the negative and zero
sequence currents also see their own network impedances.

The impedances are as seen from the fault position (the Thevenin equivalent impedances seen
upstream from the fault position).

The impedance seen by the zero sequence current (returns via the earth) is determined by the
method of earthing of the neutral of the network.

The method of earthing of the neutral of the network has no effect on the positive and negative
sequence currents (these do not flow through the earth but only between the phase conductors).

9.2.1 Balanced three-phase faults

Generators are assumed to generate only positive sequence voltages and hence under balanced
three-phase fault conditions only the positive sequence circuit has current flow.

The equivalent circuit is as follows (red phase since the red phase is the reference phase)

Z R1 I R1

VSR1 VFR1 = 0

Source Fault

Once we have solved for the above positive sequence current, all three actual phase fault currents
can be determined.

For the balanced three-phase fault condition (only the positive sequence currents are non-zero)

I R = I R1

IW = IW1 = a2 I R1

I B = I B1 = aI R1

104 ELEN5008A – Power Systems


9.2.2 Single phase-to-ground fault

The faulted circuit is as follows (the fault must be symmetrical with respect to the red phase to
correspond with our use of the red phase as the reference phase)

(a) Current phasor constraint

From the above circuit (phasor sum)

IW = IW1 + IW 2 + IW0 = a2 I R1 + aI R2 + I R0 = 0

I B = I B1 + I B2 + I B0 = aI R1 + a2 I R2 + I R0 = 0

I R = I R1 + I R2 + I R0 = I FR

This requires that I R1 , I R2 and I R0 are the same current (same magnitude and phase).

ELEN5008A - Power Systems 105


IW 1 , IW 2 and IW 0 are displaced by 120 ( I R1 , I R2 and I R0 are in phase) and since

IW1 = IW 2 = IW0

IW = 0

I B1 , I B2 and I B0 are displaced by 120 ( I R1 , I R2 and I R0 are in phase) and since

I B1 = I B2 = I B0

IB = 0

Note that even though the total currents in the white and blue phases are zero, the sequence
currents in the white and blue phases are non-zero.

Thus for the red (faulted) phase (phasor equation)

I R1 = I R2 = I R0

(b) Voltage phasor constraint

In terms of sequence voltages we need to examine the sequence voltages at the source and at the
fault.

For the red phase

Z1 Z2 Z0

VSR1 VFR1 VSR2 VFR2 VSR0 VFR0

At the fault (phasor sum) VFR = VFR1 + VFR2 + VFR0 = 0

106 ELEN5008A – Power Systems


Both the current phasor constraint and the voltage phasor constraint are satisfied with the
following sequence circuit interconnection

VSR1 VSR2 VSR0


Z1 I R1 Z2 I R2 Z0 I R0

VFR1 VFR2 VFR0

VFR

If the generator generates only positive sequence voltages (the usual assumption)

VSR1
Z1 I R1 Z2 I R2 Z0 I R0

VFR1

VFR

Once I R1 , I R2 and I R0 are solved (the reference phasor quantity is always solved first) all the
other currents can be solved

IW1 = a2 I R1

I B1 = aI R1

IW 2 = aI R1

I B2 = a 2 I R1

IW 0 = I B0 = I R0

IW = a2 I R1 + aI R1 + I R1 = 0

I B = aI R1 + a 2 I R1 + I R1 = 0

I R = I R0 + I R0 + I R0 = 3I R0 = I FR

A common mistake is to forget to multiply the calculated I R0 by three.

ELEN5008A - Power Systems 107


IB1 IW2

IR0
IR1 + I R2 + IW0
IB0

IW1 IB2

Equal magnitudes and in phase


IW = I W1 + I W2 + I W0 = 0 IR1 IR2 IR0

IB = I B1 + I B2 + I B0 = 0
IR

9.2.3 Two-phase short-circuit to ground

The faulted circuit is as follows (the fault must be symmetrical with respect to the red phase to
correspond with our use of the red phase as the reference phase)

(a) Current phasor constraint

In order to allow IR = 0

This requires (phasor sum) I R1 + I R2 + I R0 = 0

(b) Voltage phasor constraint

At the fault VFW = VFW 1 + VFW 2 + VFW 0 = 0

VFB = VFB1 + VFB2 + VFB0 = 0

VFR = VFR1 + VFR2 + VFR0  0

108 ELEN5008A – Power Systems


From (phasor sum of first two equations)

( VFW 1 + VFB1 ) + (VFW 2 + VFB2 ) + ( VFW 0 + VFB0 ) = 0

Or ( − VFR1 ) + ( − VFR2 ) + ( 2VFR0 ) = 0

Or (phasor sum) VFR1 + VFR2 = 2VFR0

Also (phasor subtraction of first two equations)( VWF 0 = VBF 0 by definition)

VFW 1 − VFB1 = VFB2 − VFW 2

This is only possible if (phasor relationship)

VFR1 = VFR2

Thus (phasor relationship) VFR1 = VFR2 = VFR0

Both the current phasor constraint and the voltage phasor constraint are satisfied with the
following sequence circuit interconnection

I R1 I R2 I R0

Z1 Z2 Z0

VFR1 VFR2 VFR0

VSR1 VSR2 VSR0

ELEN5008A - Power Systems 109


If the generator generates only positive sequence voltages (the normal assumption) the circuit
becomes

I R1 I R2 I R0

Z1

VFR1 Z 2 VFR2 Z 0 VFR0

VSR1

Once again, once I R1 , I R2 and I R0 are solved (the reference phasor quantities are always
solved first) all the other sequence and actual phase currents can be solved.

Note that the phasor diagram below is for the currents – the current phasors lag the common
voltage phasor by the respective impedance angle:

IB1

IB2
IR0
IW0
IR1 + +
IW2 IB0
IR2

IW1 IB0

IB IB2
IR1
IW1 IB1
IR0 IW
IR2
IW2
IW0
IR = 0

110 ELEN5008A – Power Systems


9.2.4 Two-phase short-circuit without-earth fault

We just set Z0 =  in the previous analysis and solve the following circuit

I R1 I R2

Z1
Z2
VFR1

VSR1

Once again, once I R1 and I R2 are solved ( I R0 = 0 )(the reference phase quantities are always
solved first) all the other sequence and actual phase currents can be solved.

IB1 IB2

IR1 + I R2

IW1 IW2
IB1
IB
IW1
I R = I R1 + I R2 = 0

IW2 IW
IB2

9.3 Concluding remarks on the above analysis

The source generally comprises known positive sequence voltages; however this does not
preclude the flow of negative sequence currents and zero sequence currents (during
asymmetrical faults).

For most network components the positive sequence impedance is equal to the negative sequence
impedance (the phase sequence does not change its behaviour)
- transmission lines
- cables
- transformers

ELEN5008A - Power Systems 111


Motors have different positive and negative sequence impedances
- for the induction motor the negative sequence impedance is obtained from the usual
equivalent circuit but setting s (slip) = 2 (rotates in opposite direction to phase sequence)
- for the synchronous motor and the synchronous generator the negative sequence impedance
is usually taken as the transient impedance of the motor or the generator

The zero sequence impedance requires special thought – it is the impedance seen looking
upstream from the fault for earth current that is shared equally between the three phase
conductors and returns via the earth.

The way that the transformer windings are connected influences the zero sequence impedance
- delta primary/delta secondary – Z0 =  (no path down to earth)
- delta primary/star (neutral earthed) secondary – Z0 no longer  (earth fault current can
flow through the star secondary winding down to earth)(the loop round the delta primary
winding provides a low impedance path (short circuit) to secondary circuit zero sequence
currents so the value of the zero sequence impedance can be quite small – leakage reactance
of transformer)
- star (neutral earthed) primary/delta secondary - Z0 =  (no path down to earth)
- star (neutral earthed) primary/star (neutral earthed) secondary – Z0 no longer  (zero-
sequence current can flow into the transformer secondary winding (earth fault on the
secondary side) and then out of the primary winding – because of mmf-balance the zero
sequence currents on the two sides of the transformer are determined by the transformer
turns ratio)

Diagrammatically (remember the sequence impedances are as seen from the fault position -
looking upstream – fault assumed on the secondary side (RHS on figures below) of transformer)

O/C Z0 Z0
  

O/C Z0 Z0

Note that even when we expect Z0 =  we still get a non-infinite value because of the stray
capacitance down to earth (can be large in cables, transformers and generators).

This current will be ignored in this discussion.

112 ELEN5008A – Power Systems


Example:

Calculate the three-phase short circuit current and the single-phase-earth fault current at the 11
kV busbar.

G T1 L T2

15 kV 132 kV 132 kV 11 kV

Rn = 1 k  R n = 18 
R n = 14,9 pu
100 MVA 100 MVA 10 MVA
Z1 = Z2 = j0,25pu Z1 = Z2 = j0,15pu Z1 = Z2 = j15  Z1 = Z2 = j0,10pu
Z 0 = j0,35pu Z 0 = j0,10pu Z 0 = j20  Z 0 = j0,10pu

Z1 = Z2 = j0,086pu Z1 = Z2 = j1,0pu
Z 0 = j0,115pu Z 0 = j1,0pu

Referring to a 100 MVA base

1322
Base impedance at 132 kV = 174,24 
100

15
Positive sequence impedance of line = 0,086 pu
174,24

For balanced three-phase short-circuit on 11 kV busbar, total positive sequence impedance


looking upstream from the 11 kV busbar is

Z1 = j( 0,25 + 0,15 + 0,086 + 1,0 )pu = j1,486 pu

Z 1(in pu) I 3(in pu)

1pu

1
Short circuit current (11 kV) I3 ( in pu ) = pu = 0,673 pu
1,486

100000
Base current at 11 kV = 5249 A
3  11

ELEN5008A - Power Systems 113


Balanced three-phase short-circuit current at 11 kV busbar

I3 ( in A ) = 0,673  5249 A = 3532A

For single-phase-earth fault on 11 kV busbar, need to determine the positive, negative and zero
sequence impedances looking upstream from the 11 kV busbar.

112
Base impedance at 11kV = 1,21 
100

18
Neutral earthing resistor in pu = 14,9 pu
1,21

1pu
Z1 (in pu) Z2 (in pu) Z 0(in pu) I (in pu)
0

Total zero sequence impedance looking upstream from the 11 kV busbar

Z0 = ( 3  14,9 + j(1,0 + 0,115 + 0,1) ) pu = ( 44,6 + j1,215 )pu

Including the positive and negative sequence impedances

Z1 + Z2 + Z0 = ( j1,486 + j1,486 + 7,5 + j1,215 ) pu = 44,6 + j4,187 pu = 44,8 pu5,36

1
I0 ( in pu ) = = 0,0223 pu
44,8

I0 ( in A ) = 0,0223  5249 = 117 A

I E / F ( in A ) = 3I0 = 351 A

114 ELEN5008A – Power Systems


CHAPTER 10: PROTECTION COORDINATION

10.1 Introduction

Protection coordination involves the management of which protection relay responds to which
fault.

Consider the following feeder circuit

CB1 CB2
Load

CB2 must respond


to fault here (and
preferably not CB1)

CB3
Load

Protection coordination can be based on


− current comparison (e.g. differential) – by measuring the current into and out of the line, can
decide whether the fault is between CB1 and CB2 so that CB1 must open
− time (e.g. IDMT) – CB1 responds slower than CB2 to allow CB2 to open first
− current magnitude (e.g. high-set overcurrent) – faults between CB1 and CB2 will generally
have higher fault current so CB1 is set to respond to these higher currents only

ELEN5008A - Power Systems 115


10.2 Comparison method – differential protection

10.2.1 Differential protection using low impedance relays

The circuit is as follows (in modern relays each CT has its own low impedance input to the relay
and the difference current is calculated using software)(the advantage of this approach is that the
CTs can be shared by multiple relays where their (low impedance) CT inputs are connected in
series)
CT1 CT2

I'1 I'2
R1 I1 R3 R 4 I2 R2
I1 - I 2

L m1 V1 Z burden V relay V2 L m2
(small)

Low impedance
Equivalent circuit of CT 1 relay Equivalent circuit of CT 2

The low impedance relay trips when the magnitude of the (spill) current it sees exceeds a preset
value (the trip value)
I1 − I 2  Itrip

A problem with the above scheme is that Current Transformers (CTs) are not ideal – in particular
two CTs of the same type do not behave in the same way so that even if the primary currents are
the same, the secondary currents could be different.

The problem is greatest when the primary currents are the greatest - during through-fault
conditions (when the relay should not operate) - the spill current due to the different behaviour of
the CTs during these high primary current conditions may be sufficient to trip the relay.

The secondary voltages (E) of the CTs determine the fluxes (  ) in the CTs

Erms = 4,44 f N  peak

At high through-fault currents, due to the resistances of the CT secondary windings, leads and
relay burdens, the voltages across the CT secondary windings ( V1 and V2 ) are high and hence
CT saturation becomes a factor that must be considered.

Small differences in the secondary voltages cause large differences in the saturation behaviour
and this causes large differences in the CT secondary currents (the CT that saturates more has a
lower magnetizing inductance in its equivalent circuit and this shunts more of the current that
would otherwise exit the CT secondary leads).

Also differences in the remnant magnetism (the flux then increases above or below this value by
the integral of the instantaneous value of the CT terminal voltage) cause differences in the instant
at which saturation of the CTs occurs.

Clearly the above circuit would be very difficult to implement reliably.

116 ELEN5008A – Power Systems


10.2.2 Differential protection using high impedance (voltage-operated) relays

High impedance differential protection tries to avoid the problems of the previous low
impedance differential protection.

The circuit is as follows (in this case the CTs are connected to a single (high impedance) input to
the relay)

CT1 CT2

I'1 I'2
R1 I1 R3 R 4 I1 R2

L m1 V1 Z burden V relay V2 L m2
(large)

High impedance
Equivalent circuit of CT 1 relay Equivalent circuit of CT 2

(a) For the through fault (relay must not operate) without CT saturation

I '1 = I '2

Ideally V1 = V2 (note the anti-parallel connection of the CTs) and the voltage across the
relay is zero and the relay does not trip the circuit breakers for the through fault (which it
should not).

(b) For the through fault (relay must not operate) with one CT saturating (the worst case
condition is when Lm1 is large ( CT1 is not saturated) and Lm2 = 0 ( CT2 is saturated so
V2 = 0 – modelled as a switch in parallel with Lm2 closing).

For this worst case condition (one CT saturated and the other not saturated) the voltage
across the high impedance relay has the value

Vrelay  I1( R2 + R4 )

Thus for stability during through-faults (relay does not trip the circuit breakers), the relay
must have a voltage setting higher than the above (generally small) value.

(c) For an internal fault (the relay must operate) with neither CT saturated

I '1  I '2

ELEN5008A - Power Systems 117


If both CT1 and CT2 do not saturate, both Lm1 and Lm2 are high and the spill current
(difference between I '1 and I '2 ) will see a high impedance ( Lm1 in parallel with Lm2 )
and the relay will see a high voltage (higher than in (b)) and the relay will trip the circuit
breakers (as it should).

(d) For an internal fault (the relay must operate) with CT saturation, as in (b) the relay is
blocked from operation if one CT saturates but this is temporary (few cycles) and
thereafter as in (c) the relay will trip the circuit breakers (as it should).

10.2.3 Differential protection using biased differential relays

The CT error problem during through-fault conditions can also be circumvented by using a
biased differential relay

I1 Low impedance Low impedance I2


restraint coil restraint coil

I1 - I2

Low impedance
operating coil

Biased differential relay

The bias coils provide a restraint such that the trip level increases with increasing through-fault
current

Operating current
I1 - I2

Large current error


causes trip Bias slope
stic
cteri
ra
p cha
Tri
Small current error
does not cause trip

Restraint current
I1 + I2

Differential protection used to protect transformers generally uses biased differential relays (tap
changer changes the transformer turns ratio while the CT ratios remain constant – introduces a
current error).

118 ELEN5008A – Power Systems


Busbar protection using differential protection generally uses high impedance relays (to cater for
saturation of the CTs).

With differential protection no time delay needs be introduced (operation can be instantaneous -
major advantage over time discrimination coordination) - except when used to protect a
transformer (to cater for the inrush current) unless second harmonic restraint is implemented).

10.2.4 Differential protection using pilot-wires (<10km)

The problem with the previous differential protection is that the full CT secondary current must
circulate (via pilot wires) between the two CTs.

With pilot wire protection, to reduce the current through the pilot wires, only the difference
current is transferred.

The circuit for the opposing voltage scheme is as follows

CT1 CT2

I1 Relay 1 Relay 2 I2

Restraint V1 V2 Restraint
coil coil

Operating Operating
coil coil

For through-faults the CT currents flow in the restraint coils of the relays and the voltages are
equal and little current flows between the relays and the relays do not operate.

For an internal fault the CT currents are unequal and hence the voltages across the restraint coils
are unequal and current flows between the relays.

The large circulating current through the operating coils operates both relays to trip the circuit
breakers at both ends.

ELEN5008A - Power Systems 119


10.2.5 Differential protection using pilot-wires and a summation transformer (<10km)

For a three phase circuit a current summation transformer is used to avoid multiple-core pilot
cables (if exact type-of-fault information is not required)

30 10 10 10 10 30

10 10
I1 Relay 1 Relay 2 I2

Restraint Restraint
coil coil

Operating Operating
coil coil

For the above current summation transformer the fault sensitivities are

Type of internal fault Relative trip settings (%)


R-GND 20 (50 primary turns)
W-GND 25 (40 primary turns)
B-GND 33 (30 primary turns)
R-W 100 (10 primary turns)
W-B 100 (10 primary turns)
B-R 50 (20 primary turns)
R-W-B 58 (phasor sum)

The higher sensitivity to earth fault currents is beneficial for high resistance earth faults.

120 ELEN5008A – Power Systems


10.2.6 Differential protection using carrier signals (> 10 km)

The two types of carrier protection are


− directional comparison
− phase comparison

Magnitudes and trip information are transferred via modulating a carrier signal (50 kHz –
500 kHz) that is injected onto the cable or transmission line.

10.2.6.1 Directional comparison

The circuit breakers at both ends are tripped when the directional relays at both ends sense that
the fault current flows into the line at both ends.

Thus when a relay at one end senses fault current flowing into the line it sends a carrier signal to
the relay at the other end.

If the relay at the same time senses a carrier signal from the other end (indicating that the relay at
the other end has also sensed a fault current flowing into the line) then the relay trips its CB.

10.2.6.2 Phase comparison

The circuit breakers at both ends are tripped when the relays at both ends see a significant
difference in phase of the currents at both ends.

Each relay modulates a carrier signal with a square wave in phase with the current it measures.

At the same time it monitors the modulated carrier signal received from the other relay.

When a significant phase shift is detected both relays trip their CB.

ELEN5008A - Power Systems 121


10.3 Time graded protection

Time graded protection becomes necessary when fault currents are monitored from one end only
- selectivity is obtained by the introduction of time delays.

Time graded protection is often used as backup protection.

The disadvantage of time graded protection is the fact that a delay to operation is added - thereby
increasing the damage at the fault location as well as increasing the stress to the upstream
equipment.

A time margin of 0,25 seconds - 0,5 seconds at maximum fault current is typical to cater for
inherent delays in the operation of the downstream protection.

Three types of time graded protection are used


− definite time overcurrent (sense current magnitude only)
− inverse time overcurrent (sense current magnitude only)
− definite time distance (sense impedance)

At the lower voltage levels ( 22 kV) time graded protection is used virtually exclusively to
protect distribution lines and network equipment.

10.3.1 Time-graded protection using definite-minimum-time overcurrent (DMT) relays

All the relays are given the same current setting (lowest fault current expected).

The operating times are as follows

1 2 3

Time to operate

1
0,5sec
2
0,5sec
3

Position of fault

Time discrimination is achieved with the penalty of large trip times for upstream faults - must
ensure that the stresses imposed on the equipment still fall within their capability.

122 ELEN5008A – Power Systems


10.3 Time grading using inverse-definite-minimum time (IDMT) overcurrent relays

The IDMT relay allows accurate coordination with the fault current withstand characteristic of
network equipment which also has an inverse characteristic due to its thermal behaviour.

I rms

Transformer withstand capability


I sc
IDMT relay characteristic

The IDMT curve allows, in one curve, fast tripping of close faults (high fault currents) and slow
tripping of remote faults (low fault currents) – allowing downstream protection time to operate.

The IDMT relay has two adjustments which allow the user to shift the IDMT curve vertically
and horizontally as required
- the current multiplier – settings typically 0,5 - 2,4 (1,0 corresponds to IN)
- the time multiplier - settings typically 0,1 - 1,0 (1,0 corresponds to the curve value)

I rms

IDMT time multiplier


IDMT relay characteristic
Downstream
protection IDMT current multiplier
Transformer rated current

The IDMT current multiplier is set such that the relay does not operate when rated current (or
some multiple, for example 1,25 times rated current) flows.

For example, if the relay has a 5 A input and is used with a CT ratio of 500:5 and 1,25 times the
rated current of the transformer being protected is 400A then this current corresponds to a relay
input current of 4 A.

ELEN5008A - Power Systems 123


Thus for the above example

4
Current multiplier = = 0,8
5

To solve for the required time multiplier we need the equation for the IDMT curve (from IEC or
IEEE standards)(standard inverse curve equation used below)(curves in data sheets correspond to
a time multiplier of 1)

0,14
t=  Time multiplier
I pu0,02 − 1

Assume coordination with downstream protection is required for a fault current of 10 kA – or

10000
I pu = = 25
400

The speed of the downstream protection is also required for this fault current of 10 kA – assume
0,7 s (perhaps from a fuse curve).

A time margin of 0,2 s is usually assumed between the upstream and downstream protection and
hence the IDMT protection must operate after 0,7 s + 0,2 s = 0,9 s for the 10 kA fault.

Thus substituting into the above equation

0,14
0,9 =  Time multiplier
0,02
25 −1

250,02 − 1
Thus Time multiplier = 0,9  = 0,427
0,14

This calculation is usually repeated for different fault currents to ensure coordination is
maintained at all possible fault currents.

124 ELEN5008A – Power Systems


When coordinating with other IDMT relays the operating times are coordinated as follows

1 2 3

Time to operate

Level of fault current

0,5sec 0,5sec
1 2 3

Position of fault

The inverse characteristic allows the use of shorter trip times for upstream faults - less stress on
upstream equipment and shorter dips seen by customers.

The inverse time characteristic is a consequence of the fault current decreasing (and hence the
trip time increasing) with distance from the respective relay.

The inverse time characteristic is inherent with


− fuses
− thermal CBs (bimetallic strip) - low cost
− magnetic CBs (with dashpot (motor contactors)) - low cost
− IDMT relays (rotating disc)
− IDMT relays (electronic)

Miniature Circuit Breakers (MCBs) and fuses have band characteristics to allow for
manufacturing spread, while overcurrent relays have line characteristics (more accurate - higher
cost).

The drawback with MCBs and fuses is that their short circuit trip characteristics are not
adjustable - problem when fault level changes (multiple generator installations).

ELEN5008A - Power Systems 125


The IDMT relay characteristics available are ( I is the pu current)

− long time delay (LTD) curve (earth fault protection of feeders)

120
t
I −1

− standard inverse (SI) curve (transformer protection)

0,14
t
I 0,02 − 1

− very inverse (VI) curve (coordination with downstream fuses)

13,5
t
I −1

− extreme inverse (EI) curve ( I 2t  constant)(coordination with downstream fuses)

80
t
2
I −1

126 ELEN5008A – Power Systems


The above curves can be plotted as follows (assuming time multiplier set at 1.0)

ELEN5008A - Power Systems 127


In the overlap regions (e.g. for relays 2 and 3) the inverse time characteristics must be
coordinated as follows

Minimum fault current


log t
seen by relay 3
Maximum fault current
seen by relay 3

Upstream
relay 2
IDMT curves
0,5sec Downstream
relay 3
Fault currents
in overlap
log I

For fuses, the upstream pre-arcing I 2t must exceed the downstream total I 2t (pre-arcing +
arcing).

The inverse time characteristic is ideal for induction motor protection where short duration high
level starting currents are expected and must not cause the protection to operate

log t
Full
speed
(normal
current)
Maximum
allowable
locked rotor
time
Acceleration
(few seconds) IDMT curve

Inrush (2 cycles)
log I

128 ELEN5008A – Power Systems


10.3.3 Time grading using definite time distance relays (reduced additional delay compared with
IDMT)

Distance protection (using impedance relays) is an example of this type of protection – except
zone division is by fault impedance rather than current.

The operating times are set as follows

1 2 3

Time to operate

0,5sec 0,5sec
1 2 3

Position of fault

Time discrimination is still present, but the distance is divided into zones with constant trip times
within these zones.

10.4 Backup protection

This operates should the main (primary) protection fail to operate within a given time.

Overcurrent (IDMT) relays tend to be used to provide this backup protection function.

ELEN5008A - Power Systems 129


10.5 Protection performance from a power quality point of view

In order to shorten the duration of voltage dips that other customers are exposed to, the
protection should clear the fault as quickly as possible

log t

Get as close as possible

IDMT curve
Fuse
characteristic

log I

In some critical situations, more expensive protection (e.g. instantaneous differential) replaces
cheaper IDMT relays if power quality is a major consideration.

Customer equipment has the following sensitivity to voltage dips due to remote faults – reason
why want to reduce operating times of protection relays:

Residual voltage
Normal operating voltage (nominal)
Equipment
survives

Equipment
Better maloperates or trips

Dip duration
Typically
10's of ms

130 ELEN5008A – Power Systems


CHAPTER 11: TRANSFORMER PROTECTION

11.1 Introduction

Types of protection used with transformers include


− Buchholz – float device that can detect oil bubbles and oil turbulence caused by an internal
transformer fault – placed in the oil pipe between the main transformer tank and the
overhead oil conservator
− tank overpressure – detect internal faults
− oil temperature monitoring device (lower threshold for alarm purposes and higher threshold
for trip) – protect against transformer overloads
− winding temperature monitoring device (lower threshold for alarm purposes and higher
threshold for trip) – protect against transformer overloads
− inverse-definite-minimum-time (IDMT) overcurrent relay on the primary side – protect
against transformer overloads and internal transformer faults, and provide backup protection
for downstream faults
− instantaneous differential – protect against primary and secondary winding faults
− earth fault relay – protect against primary and secondary faults to earth and backup
protection for downstream earth faults
− restricted earth fault (REF) relay on the primary winding – protect against primary winding
faults to earth
− restricted earth fault (REF) relay on the secondary winding – protect against secondary
winding faults to earth

ELEN5008A - Power Systems 131


11.2 IDMT relay on the primary side (both small and large transformers)

This has the connection (star or delta primary winding)

IDMT
relay

This IDMT relay provides


− protection for transformer overloads
− protection for internal transformer faults
− main protection for small transformers
− backup protection for large transformers where the main protection is differential
− backup protection for downstream faults

The IDMT relay usually also has an instantaneous high-set characteristic that operates
immediately if the current exceeds that for a solid three phase short circuit on the secondary side
(this implies a fault internal to the transformer).

Providing backup protection for downstream faults is important because the transformer itself
can be damaged by long-duration through-faults.

For increased reliability, backup protection should use its own dedicated CTs.

132 ELEN5008A – Power Systems


11.3 Differential protection (larger transformers)

For a star-star transformer (CTs must be delta connected to avoid zero sequence currents due to
external earth fault causing trips)

Differential
relay

1 1
Ratio of CT turns ratios = =
Transformer turns ratio NT

For a delta-star transformer (very common for MV/LV)

Dyn1
11kV 420V
I R = I RW - IBR 200:1 5400:0,577 Ir
IRW BR
IBR r
RW
b w
I W = I WB - IRW WB Iw
IWB
I B = I BR - IWB Ib

iB iw - i b iw ir ib

iW i b - ir

iR ir - iw

Differential
relay
CTs star-connected CTs delta-connected

ELEN5008A - Power Systems 133


IR Ir
iR = =
NCTprim NCTprim NT

3I r
ir − iw =
NCT sec

Thus, we need to set NCT sec = 3NCTprim NT

For example, for an 11 kV/408 V Dyn1 transformer

11000
NT = = 27
408

And if NCTprim = 200

Then NCT sec = 3  200  27 = 3  5400

Usually expressed as 5400 : 0,577

For a star-delta transformer


YNd1
IR 100:0.577 132 kV 22 kV 600:1 I r = I rw - I br
br I rw
R I br
rw
W B
IW wb I w = I wb - I rw
I wb
IB I b = I br - I wb

iW iR iB iW - i B ib

iB - iR iw

iR - i W ir

Differential
relay
CTs delta-connected CTs star-connected

Connecting the CTs in a delta on the star side ensures that zero sequence currents do not escape
into the differential relay even when there is significant neutral current (external earth fault).

Differential protection provides protection for primary and secondary winding faults.

Operation can be instantaneous (no added delay).

134 ELEN5008A – Power Systems


Because large transformers inevitably have tap changers (turns ratio changers) the relay setting
must be high enough so that there is no spurious operation for through-faults when operating on
any of the taps.

Typical characteristic (biased differential)

For transformer protection, a second harmonic restraint characteristic is usually included to block
tripping due to the large inrush currents associated with transformer energization (inrush current
has a distinctive second harmonic current component).

11.4 Earth fault relay (smaller transformers)

By measuring the current between the neural point of a star winding and earth, an earth fault can
be detected either in the respective transformer winding or in the external network

Adding an IDMT characteristic allows


− backup for other downstream earth fault protection relays
− backup for REF protection (for large transformers the main earth fault protection is REF)

ELEN5008A - Power Systems 135


11.5 Restricted earth fault (REF) and balanced earth fault (BEF) relays (larger transformers)

That the above earth fault relay also responds to earth faults outside the transformer can be a
disadvantage. The restricted earth fault and balanced earth fault relays only respond to earth
faults inside the transformer.

The circuit is shown below (star winding with earthed neutral)

For a delta winding (or star winding but with an unearthed neutral)

Provided the CTs have the same ratio, an external earth fault does not cause current through the
earth fault relay.

The magnitude of the earth fault current when the REF or BEF is on the primary is determined
by how the transformer supplying the above transformer is earthed
– high if secondary winding star connected with neutral solidly earthed (set at 0,2pu)
– low if secondary winding delta or star connected with neutral earthed via impedance

136 ELEN5008A – Power Systems


When setting up REF and BEF relays, the capacitive current to earth needs to be considered.

Under healthy network conditions, the phase-earth capacitance currents are equal and displaced
by 120 and hence do not cause a non-zero residual current in the REF and BEF circuits.

However if an earth fault occurs outside the zone of protection of the REF or BEF relays this
causes an unbalance of the capacitive current inside the zone of protection because the one phase
voltage is zero and the healthy phase voltages rise to up to 3 times their normal value.

Thus there is residual capacitive current (note the absence of the red phase capacitive current
because the red phase voltage is assumed to be zero and the voltage to earth of the healthy
phases is now 3 V and not V )

Typically, the cable capacitive current under healthy conditions is quoted (in A/km) and hence to
avoid spurious operation of the REF or BEF relay for out-of-zone earth faults

I setting = 3  Normal phase-to-earth capacitive current

- important when a long cable length is inside the REF or BEF zone of protection (significant
capacitive currents).

ELEN5008A - Power Systems 137


11.6 Example of calculation for transformer IDMT protection (IDMT relay on primary side of
transformer)

Consider a 132 kV/22 kV (star-delta) 50 MVA transformer with an 8% impedance.

Assume the 132 kV fault level is 1000 MVA (5% source impedance on 50 MVA (transformer
MVA rating) base).

50000 50000
Transformer rated current = = 219 A (132 kV side) and = 1312 A (22 kV
3  132 3  22
side).

Suitable CT ratios = 300/1 (132 kV side) and 2000/1 (22 kV side).

1,25  219 1,25  1314


Current multiplier = = 0,913 (132 kV side) and = = 0,82 (22 kV side).
300 2000

For a fault level of 1000 MVA on the 132 kV side of the transformer, the short circuit current
seen by the IDMT relay on the 132 kV side of the transformer will be

1000 MVA
= 4374 A
3  132 kV

4374
In per unit (132 kV side): I pu = = 16
1,25  219

Assume that the IDMT relay must operate in 0,2 second for a solid three-phase fault on the 132
kV side of the transformer.

The IDMT settings must therefore be (SI curve)

- current multiplier (132 kV side)

0,913

- time multiplier (to give the required trip time of 0,2 sec)

0,14
0,2 =  Time multiplier
0,02
16 −1

16 0,02 − 1
Time multiplier = 0,2  = 0,08
0,14

138 ELEN5008A – Power Systems


CHAPTER 12: TRANSMISSION LINE PROTECTION

12.1 Introduction

Transmission line protection can have the form


- unit (main protection)
- differential pilot wire
- phase comparison carrier
- non-unit (backup protection)
- overcurrent (e.g. IDMT)
- distance (impedance))

The protection schemes used for distribution lines ( 132 kV) are
- definite time overcurrent (cheapest and slowest)
- inverse time overcurrent
- directional inverse time overcurrent
- differential pilot wire (most expensive but fastest)

The protection schemes used for transmission lines (> 132 kV) are
- distance (impedance)(very common)
- differential pilot wire or phase comparison carrier (most expensive but fastest)

ELEN5008A - Power Systems 139


12.2 Directional overcurrent protection

The overcurrent trip is blocked if the direction of the (fault) power flow is not in the direction the
directional relay is looking – the relay is enabled only if

−90   −   90

where  is the preset characteristic angle of the relay


 is the measured impedance angle of the fault current

Typically  = 60 for a line fault thus set  = 0 if measure line current and phase-neutral
voltage, or  = 30 if measure line current and line voltage (the 30 is to compensate for the
inherent phase angle between the line and phase-neutral voltages).

jX

Downstream fault
(relay enabled)

Upstream fault Characteristic angle
(relay blocked)
R of relay,  = 0º

jX
Downstream fault
(relay enabled)
Characteristic angle
of relay,  = 30º


R
Upstream fault
(relay blocked)

140 ELEN5008A – Power Systems


Directional protection is applied where parallel supply paths are present

The protection is coordinated in two groups


- ABCD – for fault between Q and D, D must trip before A so that both feeders are not lost
- QRST – for fault between A and T, T must trip before Q so that both feeders are not lost

ELEN5008A - Power Systems 141


12.3 Distance (impedance) protection

Distance protection is essentially a non-unit protection providing both main and backup
protection functions within a single relay.

Distance protection monitors the ratio between the voltage and the current – the impedance seen
by the relay.

The measured current can be considered as the operating quantity and the measured voltage as
the restraining quantity.

There is an impedance measuring module for each phase-to-phase and each phase-to-ground
combination.

The respective quantities are as follows

Type of fault Vrestraining I operating


R-W VRW I R − IW
W-B VWB IW − I B
B-R VBR IB − IR
R-G V RG I R + kI N
W-G VWG IW + kI N
B-G V BG I B + kI N

The distance relay measures all currents and voltages and decides on the type of fault (which
voltages and currents to use in the respective impedance calculation).

During a fault, the impedance seen by the relay drops dramatically – the point on the impedance
curve moves towards the origin.

The simple constant impedance magnitude locus is as follows (trip initiated if measured
impedance moves into circle)

142 ELEN5008A – Power Systems


The simple constant impedance magnitude characteristic has a significant disadvantage - the
distance protection would sense a fault on both sides of the relay (provided there is a source of
fault current in the opposite direction – perhaps the line being protected is one of two parallel
lines) – a directional characteristic is preferred.

This can be achieved by allowing operation only for positive resistances and positive reactances
(as seen by the distance relay)

If we consider the range of fault impedances seen by the relay as well as the normal load
impedance

The impedance relay settings must be adjusted according to the CT and VT turns ratios

CT ratio
Secondary ohms = Primary ohms 
VT ratio

ELEN5008A - Power Systems 143


A relay with the trip characteristic below (approximation of the above) is known as a mho relay
and is very widely used

The name given to the above characteristic (mho) is because the above circle in the impedance
(ohm) plane maps to a straight line in the admittance (mho) plane.

If we assume the center of the circle is at point (a,b) and the circle passes through the origin, then
the equation for the circle is

( R − a )2 + ( X − b )2 = a 2 + b2

Or R2 − 2aR + X 2 − 2bX = 0

2aR 2bX
Or + =1
R2 + X 2 R2 + X 2

But if we rather measure the admittance

I 1 R X
Y= = G + jB = = −j
V R + jX R + X
2 2
R + X2
2

Substituting into the above 2aG − 2bB = 1

144 ELEN5008A – Power Systems


Which can be written in the form ( y = mx + c )

a 1
B= G−
b 2b

- the straight line in the admittance plane

Other types of characteristics are shown below:

Cross-polarized Mho

ELEN5008A - Power Systems 145


Offset Mho (center of circle along transmission line impedance locus)

Lens (Lenticular)

Quadrilateral (also very popular)

146 ELEN5008A – Power Systems


The problem of load encroachment (distance protection mistaking a low load impedance
(overload condition) for a line fault, is tackled using a load encroachment blinder (impedances
within the blinder area do not cause distance relay operation)

The load encroachment blinder usually only comes into effect if the voltage is above 0,7pu – a
lower voltage implies that a fault condition does in fact exist.

Incorrect distance protection operation during short-duration overload conditions has been
blamed for many recent well-publicized wide-area blackouts.

ELEN5008A - Power Systems 147


12.2.1 The implementation of zoning with distance protection

Consider the following transmission lines with intermediate busbars (substations) B and C

The distance relays looking into the respective transmission lines have the following zoning
characteristic

The zone 1 element can operate within 1 cycle (20ms) of the fault occurring.

The zone 2 element is deliberately delayed to allow downstream protection to operate first
(provides backup for the downstream protection) – similarly for the zone 3 element.

148 ELEN5008A – Power Systems


The distance relay at circuit breaker A has the following set of zone characteristics

There is a set of the above characteristics for each phase-to-phase and each phase-to-ground
combination.

The zone 3 element of relay A is sometimes given a small reverse reach (typically 20% of the
protected line) in addition to the forward reach.

This provides time delayed local backup protection for upstream busbar faults and close-up
downstream three phase faults (the zone 1 protection may not be accurate for very close faults).

The zone 3 element is often set to an impedance corresponding to where the transmission line
operates at 150% of its thermal limit (well clear of overload conditions).

With carrier accelerated distance protection for a fault in the final 20% of the transmission line,
because it is within zone 1 of the remote distance protection relay the remote relay uses a carrier
signal injected onto the transmission line to instantaneously trip the circuit breakers at both ends
(now distance protection acts as unit protection).

The distance protection may also provide backup protection to other unit protection schemes
- differential pilot wire main + distance backup
- phase comparison carrier main + distance backup

Similarly, where distance protection is the main protection, the backup protection may be IDMT
overcurrent protection.

ELEN5008A - Power Systems 149


12.2.2 Out-of-step blocking of distance protection

Out-of-step blocking is required when there are voltage sources (generators) at both ends of the
transmission line.

Out-of-step blocking is required to prevent the distance protection operating when then is no
fault along the transmission line – there is just a loss of synchronism between the generators.

During an out-of-step condition the voltage phasors at the ends of a transmission line rotate with
respect to each other

The current along the transmission line therefore varies with V and is a maximum for
 = 180 and a minimum for  = 0 .

Clearly the impedance measured at the one end, V/I, varies between a small value (  = 180 )
and a large value (  = 0 ).

When  = 180 the impedance relays see a short circuit at the center (electrical center) of the
line.

The impedance loci traced during an out-of-step condition are as follows

150 ELEN5008A – Power Systems


VA
- where k=
VB

2
V
At  = 180 or 0 , P = 0 and hence Rmeasured = P =0
2 2
P +Q

At  = 180 , Q is a maximum and hence X measured is a minimum.

At  = 0 , Q is a minimum and hence X measured is a maximum.

Thus during the slip cycle for  = 180 the impedance locus enters the trip areas of the
impedance relays.

To avoid tripping, an out-of-step (OOS) blocking zone is added

If the out-of-step (OOS) zone is entered (impedance always starts high -  is small) and there is
a significant delay before the zone 3 is entered (because of the low slip frequency), then tripping
is blocked (in contrast, with a true transmission line fault this transfer occurs very quickly and
tripping is enabled).

ELEN5008A - Power Systems 151


12.3.3 Measurement of impedance for distance protection

The positive sequence impedance, Z1 , must be accurately measured for all types of faults.

(a) Phase-phase, three-phase and two-phase-to-earth faults

For a phase-phase fault involving the white and blue phases must measure at the measurement
location

VWB and IW − I B

I R1 I R2

Z1 V'R1
Z2 VR2

VMR1

At the measurement position ( I R1 and I R2 are in antiphase)(assuming Z1 equals Z 2 )( Z1 and


Z 2 are now up to the fault position)

VWB = VW − VB = VW 1 + VW 2 + VW 0 − VB1 − VB2 − VB0

2 ' 2
= (a − a)VR1 + (a − a )VR 2

2 2
= (a − a) I R1Z1 + (a − a) I R1Z1

2
= 2(a − a) I R1Z1

For the currents IW − I B = IW 1 + IW 2 + IW 0 − I B1 − I B2 − I B0

= (a2 − a) I R1 + (a − a 2 ) I R 2

= (a2 − a) I R1 + (a2 − a) I R1

= 2(a2 − a) I R1

152 ELEN5008A – Power Systems


2
VWB 2(a − a) I R1Z1
Therefore = = Z1
IW − I B 2(a 2 − a) I R1

(b) Earth fault

For an earth fault involving the red phase must measure at the measurement location

VRG and I R + kI N

VRG
Z1 I R1 Z2 I R2 Z0 I R0

- where I N is the residual current (= 3I0 ), and the residual (neutral) or ground compensation
factor, k , is
Z − Z1
k= 0
3Z1

- which allows the measurement of the positive sequence impedance to the fault position.

Thus (for an earth fault the sequence component currents are in phase)(assuming Z1 equals Z 2 )
( Z1 , Z 2 and Z0 are from the measurement position up to the fault position)

VRG = I R1Z1 + I R2 Z 2 + I R0 Z0

= I R1Z1 + I R2 Z1 + I R0 Z1 + I R0 ( Z0 − Z1 )

= I R Z1 + I R0 ( Z0 − Z1 )

  Z − Z1 
=  I R + I R0  0   Z1
  Z 
  1 

ELEN5008A - Power Systems 153


 I  Z − Z1 
=  IR + N  0   Z1

 3  Z1 

VRG
Or Z1 =
I R + kI N

154 ELEN5008A – Power Systems


CHAPTER 13: POWER SYSTEM ECONOMICS

13.1 Concept of the capacity factor

The capacity factor is a measure of the energy (MWh) that can be expected from a power plant
relative to its rated capacity (MW) multiplied by the time (typically one year – 8760 hours).

Energy produced in one year ( MWh )


Capacity factor =
MWrated  8760

Capacity factors tend to be low for renewable energy sources as they are weather dependent
(wind is not always blowing or sun is not always shining).

The capacity factor plays a major role in indicating the required tariff that a power plant must
receive to be financially viable.

13.2 Concept of the Levelized Cost of Energy (LCOE)

The Levelized Cost of Energy ( LCOE ) allows comparison of costs between different power
plant technologies.

Just as the power plant produces income for every year of operation, we need to balance this
against the different forms of costs for every year of operation.

Calculating the annualized capital cost is where the capital to build the power plant is converted
into annual repayments (annualized capital cost) where the fractional interest paid on the debt
per year is given as r and the duration of the loan is given as n years.

Assume the initial (start of year one – present day) capital expenditure on the power plant is $K
million and the annual repayment of the loan (paid at the end of each year) is $A million.

$ K million( 1 + r )n = $A million(1 + r )n−1 + ... + $ A million( 1 + r )1 + $ A million

At start of year 1, At end of year 1, At end of year n-1 At end of year n,


repay $K million – repay $A million – repay $A million – repay $A million
avoid repayment of avoid repayment of avoid repayment of
$K million with $A million with $A million with
interest incurred interest incurred over interest incurred
over n years n-1 years over 1 year

The above equation is usually written in the following form (each term now represents the
present-day value – the effective value of $A million repaid in the future has been reduced
because of it being held so long and therefore incurring interest)

$ A million $ A million $ A million $ A million $ A million


$K million = + + + + ... +
(1+ r ) (1+ r )
2
(1+ r )
3
(1+ r )
4
(1+ r )
n

ELEN5008A - Power Systems 155


Multiplying all terms in the equation by ( 1 + r )−1 gives

$K million $ A million $ A million $ A million $ A million $ A million


= + + + + ... +
(1+ r ) 2 3 4 4 n +1
(1+ r ) (1+ r ) (1+ r ) (1+ r ) (1+ r )

Subtracting the second equation from the first equation gives

$K million $ A million $ A million


$K million − = −
(1+ r ) (1+ r ) n +1
(1+ r )

Multiplying all terms by ( 1 + r )n +1 gives

$K million(1 + r )n+1 − $K million(1 + r )n = $A million(1 + r )n − $ A million

But (1 + r )n+1 − (1 + r )n = (1 + r − 1)(1 + r )n = r(1 + r )n

Thus (known as the annualized capital cost (or annuity equation))

r( 1 + r )n $K million
$ A million = $K million and not
n
(1+ r ) −1 n

To this is added the annual additional build cost to get the total annualized build cost, I t .

The annual Operations and Maintenance (O&M) cost, M t , and the annual fuel cost, Ft , are
added to get the cost to be recovered every year (year t ) expressed as its present-day value

It + M t + Ft
t
(1+ r )

Capacity factor (%)


- where Ft =  MWrated  8760  ($ / MWh ) fuel,t
100

Then dividing by the annual generated energy (kWh) gives the equivalent tariff required to
recover the costs for the assumed capacity factor (Levelized Cost of Energy ( LCOE ))

Sum of costs (present − day values) over lifetime


LCOE =
Sum of energy ( present − day values ) produced over lifetime

156 ELEN5008A – Power Systems


n I +M +F
 t t
t
t
(1+ r )
= t =1
n Et
 t
t =1 ( 1 + r )

- where I t is the total annualized build cost in year t (in $)


M t is the total Operations and Maintenance (O&M) cost in year t (in $)

M t = M fixed ,t + M variable,t = M fixed ,t + Et  ($ / MWh )O&M ,t

Ft is the fuel cost in year t (in $)

Et is the electrical energy generated in year t (in MWh)(also divided by ( 1 + r )t


because energy produced in year t has lost some of its value compared to if it
had been produced in the first year of operation)

Et ( in MWh ) = Capacity factor  MWrated  8760

r is the Weighted Average Cost of Capital (WACC)(diversified ways of raising


capital – shares (equity) and bank debt)(as a fraction)
n is the expected economic lifetime of the power plant (in years)

For example (US Energy Information Administration (EIA), February 2021), LCOEs for
selected technologies (coming into service in 2026)(in 2020 US dollars)

Power plant Capacity Levelized Fixed Variable Transmission LCOE


type Factor Capital O&M Cost Investment ($/MWh)
Cost Cost (including Cost
($/MWh) ($/MWh) fuel) ($/MW)
($/MWh)
Coal, ultra- 85% 43,80 5,48 22,48 1,03 72,78
supercritical
Advanced 90% 50,51 15,51 9,87 0,99 76,88
nuclear
Geothermal 90% 19,03 14,92 1,17 1,28 36,40
Biomass 83% 34,96 17,38 35,78 1,09 89.21
Wind, onshore 41% 27,01 7,47 0.0 2,44 36,93
Wind, offshore 44% 89,20 28,96 0.0 2,35 120,52
Solar PV 29% 23,52 6,07 0.0 3,19 32,78
Solar PV with 28% 31,13 13,25 0,0 3,29 47,67
battery storage
Hydroelectric 55% 38,62 11,23 3,58 1,84 55,26

ELEN5008A - Power Systems 157


13.3 Process and financing of non-Eskom (Independent Power Producer (IPP)) renewable power
plants (e.g. wind farms and solar PV parks) in South Africa

Two approaches to financing are shown below – Special Purpose Vehicle (SPV) or developer
financed:

DMRE: Department of Mineral


DMRE offers X MW in Bidding Window n Resources and Energy (RSA)

IPP: Independent Power Producer

IPP conducts due diligence EIA: Environmental Impact


(feasibility study)(commence design) Assessment
(EIA)(Eskom connection quotes)
PPA: Power Purchase Agreement

SPV: Special Purpose Vehicle


(separated legal entity)
IPP raises funding / agrees on funding
EPC: Engineering, Procurement
and Construction

IDC: Industrial Development


IPP bids Y MW @ A c/kWh Corporation (RSA)
and other socio-economic commitments

Other

Banks lend money IPP proclaimed 'preferred bidder' by DMRE


(for interest)(debt) Other

Banks buy equity Either ... ... or Banks lend money


(for dividends) (for interest)(debt)
SPV IPP finalises capital Developer
Investors buy equity Banks buy equity
(for dividends) (for dividends)

IDC buys equity Investors buy equity


(for dividends) IPP appoints EPC contractor (for dividends)

Developer buys equity IDC buys equity


(for dividends) (for dividends)
(limits its potential
loss) IPP reaches financial close

Debt is cheaper than equity


(otherwise potential
shareholders would just IPP declared bid winner by DMRE
put their money into bank
savings accounts)

Debt:equity ratios much higher


than that of corporate debt PPA with Eskom @ A c/kWh
due to lower risk as seen by
banks

Mainly non-recourse funding


Construction starts

Private equity: equity in a non-public company (not listed on a stock exchange)(players are institutional investors, pension funds)

Public equity: equity in a company listed on a stock exchange

Private equity players have greater control over the company (have longer investment horizon prior to exiting at the optimum time)

158 ELEN5008A – Power Systems


13.4 Concept of the Net Present-day Value (NPV) for investment decision making

To calculate Cash Flow ( CF ) for year t

CFt = ( income − costs )t = ( Et  tariff t ) − ( I t + M t + Ft )

CFt
- with present-day value
t
(1+ r )

Summing for all the years of operation we get the Net Present-day Value ( NPV )

n CFt
NPV = 
t
t =1 ( 1 + r )

Decision is: if NPV > acceptable value then invest


if NPV < acceptable value then do not invest

The value of r is used where there is no risk (equivalent to interest earned in a fixed-interest
bank account).

The Discount Rate, R , is used to discount (hence the name – $1 earned tomorrow is worth less
than $1 earned today) future cash flows into present-day values where risks are now present
(more realistic)

Rr

13.5 Concept of the Internal Rate of Return (IRR) for investment decision making

In this case the company decides what (Internal) Rate of Return ( IRR ) is acceptable to itself
(reason why referred to as internal)(e.g. to pay dividends to its equity partners).

Then we revise the above equation to (note the zero on the LHS of the equation or NPV = 0 )

n CFt
0 =
t
t =1 ( 1 + IRR )

Decision is: if IRR acceptable then invest


if IRR not acceptable then do not invest

The acceptable IRR would be a value larger than the Discount Rate, R .

The NPV approach is more reliable if the cash flow jumps between positive and negative
(additional capital invested) values during the lifetime of the power plant (have multiple
solutions (values of IRR ) for the above – Descartes’ rule).
ELEN5008A - Power Systems 159
13.6 Cost of Unserved Energy (COUE)

The COUE is used for decision making – is it financially justifiable to spend money on power
system infrastructure (network and / or generation) to make the customer supply more reliable
(fewer and shorter duration interruptions)?

This is often referred to as Value Based Reliability Planning ( VBRP ) – find an optimal
compromise between network and / or generation expenditure and reliability improvement.

The difficulty is that the Customer Interruption Cost ( CIC ) is dependent on many factors
- Temporal (time of day, season, whether notice was given, number of times (frequency))
- Spatial (region)
- Customer-related (dependency on electricity, resilience to interruptions (owns UPS))

13.6.1 Micro-economic approach to CIC

In this approach the cost of interruptions is obtained from considering the actual customers.

A typical equation for the Customer Interruption Cost ( CIC ) could have the form

CIC =    ( Ni, j Fi, j,k Ci, j,kVi, j,k )


i j k

- where i is the customer class


j is the region where the customer is located
k is the reliability event type (e.g. interruptions of duration 1 - 10 minutes)
N i, j is the number of customers of class i in region j
Fi, j,k is the frequency (perhaps number per year) of reliability event of type k
experienced by customers of class i in region j
Ci, j,k is the cost to customers of class i in region j when experiencing a reliability
event of type k
Vi, j,k is the vulnerability of customers of class i in region j to reliability events of
type k (has value between 0 and 1)

13.6.2 Macro-economic approach to CIC

This is the simplest approach since it aggregates customers and uses easily obtained macro-
economic data – e.g. Gross Domestic Product ( GDP )

Gross Domestic Product (GDP)of the whole country(in R/year)


COUE (in R/kWh) =
Total energy consumption of the whole country(in kWh/year)

The first problem with the above equation is that the GDP includes taxes and excludes subsidies
so we rather replace the GDP with the Gross Value Added ( GVA ) where

GVA = GDP − Taxes ( on products only − not salaries ) + Subsidies ( on products only )

160 ELEN5008A – Power Systems


GrossValue Added (GVA) of the whole country(in R/year)
COUE (in R/kWh) =
Total energy consumption of the whole country(in kWh/year)

The second (more serious) problem is that it is too broad – it looks at the whole economy.

A preferred approach is to break the economy into economic sectors.

The annual GVA per economic sector is readily available (www.statssa.gov.za) and the annual
energy consumption per economic sector is also readily available either from Statistics SA or
from Eskom.

The equation then becomes

GrossValue Added (GVA)of a particular economic sector (in R/year)


COUE (in R/kWh) =
Total energy consumption of a particular economic sector (in kWh/year)

The third problem is that the above equation assumes each economic sector operates in isolation
(known as the direct COUE ), whereas an interruption in one economic sector also has a ripple
effect on other economic sectors (known as the indirect COUE ).

A method for calculating the total effect (both direct and indirect) was approved by the National
Energy Regulator of South Africa (NERSA) in 2016 (www.nersa.org.za).

This method (involves multiplying matrices by vectors) is outside the scope of this course, but
typical results (for 2015) are shown below

ELEN5008A - Power Systems 161


Economic sector COUE (direct effect only - ignores ripple effects) COUE (both direct and indirect effects)

Agriculture R11.55 /kWh R42.21 /kWh

Mining R13.06 /kWh R54.05 /kWh

Manufacturing R5.84 /kWh R54.64 /kWh

Electricity and water supply R7.70 /kWh R29.31 /kWh

Construction R204.10 /kWh R385.55 /kWh

Trade R108.65 /kWh R136.90 /kWh

Transport and communication R87.29 /kWh R348.64 /kWh

Finance R105.77 /kWh R400.22 /kWh

Community services R159.39 /kWh R319.37 /kWh

General government R66.62 /kWh R80.33 /kWh

Total economy (2015) R23.81 /kWh R84.16 /kWh

This should be compared with the household COUE of R6.77 /kWh.

162 ELEN5008A – Power Systems

You might also like