Sie sind auf Seite 1von 47

SYSTEM STABILITY UNDER

POWER ELECTRONIC
CONTROL

8.1 Introduction
The presence of high voltage power electronic equipment in the transmission system
has important consequences on the stability of the power system. However, the conven-
tional multi-machine stability assessment described in the last chapter is inadequate to
represent the transient behaviour of some power electronic devices. This problem is
particularly important in the case of HVDC converters due to the occurrence of inverter
commutation failures that often follow a large disturbance. Only the use of electro-
magnetic transient programs, described in Chapter 6, can provide accurate information
of the power electronic response during and immediately following clearance of the
disturbance. Although machine rotor dynamics, as discussed in Chapter 7, could be
included in the electromagnetic transient program of Chapter 6 , this is not a practical
proposition in the presence of power electronic switching, considering the different
time constants involved. For instance, electromagnetic-transient simulations require
steps of (typically) 50 ps, whereas the stability programs use steps at least 200 times
larger.
To reduce the computational requirements, some programs [ 11 contain two separate
modes. An instantaneous mode is used to model components in three-phase detail with
small time steps in a similar way to the EMTPEMTDC programs [2]. The alternative
is a stability mode based on r.m.s. quantities at fundamental frequency only, with
increased time-step lengths. The user can switch between the two modes as required
while running but, in either mode, the entire system must be modelled in the same
way. Thus, when using the instantaneous mode, a system of any substantial size would
still be very computationally intensive.
A more efficient solution is described here [3,4] which takes advantage of the
computationally inexpensive dynamic representation of the a.c. system in the stability
program, and the accurate transient simulation of the power electronic devices.
The slow dynamics of the a.c. system are adequately modelled by the stability
program while the fast dynamic responses of the HVDC and FACTS devices are accu-
rately represented by electromagnetic transient simulation. This hybrid combination is
essential to predict the first few swings in stability studies.
298 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

Finally, quasi steady-state models of the power electronic devices are also developed
in this chapter for use in longer stability studies as well as system response to small
perturbations.

8.2 Description of the Algorithm


Transient stability (TS) studies are normally carried out on the assumptions of balanced
and sinusoidal waveforms and such assumptions need to be extended to the voltage
and current waveforms at the terminals of the power electronic devices. As explained
in the introduction, part of the study requires detailed transient simulation, the results
of which cannot be directly incorporated in a TS study.
It is therefore necessary to extract fundamental frequency quantities for the TS study
from the voltage and current waveforms produced by the electromagnetic transient
program (which in this book is the PSCADEMTDC version).
The hybrid concept, referred to as TSE and illustrated in Figure 8.1, uses electrome-
chanical simulation as the steering program, and the electromagnetic transient program
is called as a subroutine. The interfacing code is written in separate routines to mini-
mize the number of modifications and thus make it easily applicable to any stability
and dynamic-simulation programs.
Initially, the TSE hybrid reads in the data files and runs the entire network in
the stability program, until electromechanical steady state equilibrium is reached.
The steady-state representation of the converter, described in Chapter 3, is perfectly
adequate as no fault or disturbance has yet been applied. Before placing the network
disturbance, the TS network is split into two independent and isolated systems, system
1 and system 2, as shown in Figure 8.2.
System 1 contains the a s . part of the system modelled by the stability program TS,
and system 2 the part of the system modelled in detail by EMTDC.
The snapshot data file is now used to initialize the EMTDC program used, instead
of the TS representation of system 2. The two programs are then interfaced and the
network disturbance can be applied. The system 2 representation in TS is isolated but
kept up to date during the interfacing at each TS time step to allow tracking between
programs. The a.c. network of system 1 modelled in the stability program also supplies
interface data to the system 2 network in TS, as shown in the lower part of Figure 8.2.
During the disturbance, the quasi steady-state representation of system 2 in TS and
the EMTDC representation of system 2 are tracked. When both of these system 2

Conventional Stability
Analysis

7- transient simulation
of power electronic

Figure 8.1 The hybrid concept


8.2 DESCRIPTION OF THE ALGORITHM 299

Stability program ENTDC program

Figure 8.2 Interfacing procedure

detailed

sequence Stability
data for Loadflow input TSlEMTDC Outflow
whole hybrid

c-y
Interface

Figure 8.3 Data flow

models produce the same results within a predefined tolerance and over a set period,
the complete system can then be reconnected and used by TS,and the EMTDC repre-
sentation terminated. This allows better computational efficiency, particularly for long
simulation runs.

8.2.1 Data flow


Data for the EMTDC solution entered in the programme database via the PSCAD
graphics. Equivalent circuits are used at each interface point to represent the rest of
the system not included in the detailed model. This system is then run until steady
300 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

state is reached and a snapshot is taken. This snapshot holds all the relevant data for
the various system components and is used as the starting point when interfacing the
detailed model with the stability programme.
The stability program is initialized conventionally through power-flow results via a
data file. An interface data file is also read by the TSE hybrid that provides the number
and location of interface buses, analysis options, and timing information. The data flow
diagram is shown in Figure 8.3.

8.2.2 Modifications required to the component programs


The main modification of the TS program is a call of the interfacing routine in the main
loop, as shown in Figure 8.4. The complete network is split into system 1 and system
2 at the interface points, but this task is performed using separate code. The only other
direct modification inside TS is the inclusion of the interface current injections at each
TS network solution.
To enable EMTDC to be called as a subroutine from TS requires a few changes
to its structure. This is helped by splitting EMTDC into three distinct segments, i.e.
an initializing segment, the main time loop, and a termination segment. In this way,
TS can call the main time loop for discrete periods as required when interfacing. The
EMTDC options, which are normally available at the beginning of a simulation run,
are moved to the interface data file and read from there. The equivalent-circuit source
values, which TS updates periodically, are located in the user-accessible DSDYN file
of the EMTDC (described in Chapter 6).

8.3 TS/EMTDC Interface


The hybrid algorithm involves regular exchanges of information between the transient
stability and electromagnetic transient programs.
The information to be transferred from one program to the other must be sufficient
to determine the power flow in or out of the interface. As shown in Figure 8.5, possible
parameters to be used are the real power, P, the reactive power, Q, the voltage, V, and
the current, I, at the interface. Phase-angle information is also required to be able to
use separate phase frames of reference.
Appropriate equivalent circuits are needed to represent the network modelled in the
stability program when running the EMTDC and vice versa. These are illustrated in
Figure 8.6, where E, and represent the equivalent circuit of system 1 and 7, and
z2 the equivalent circuit of system 2.
8.3.1 Equivalent circuit components
In Figure 8.6, f, and z2 represent the detailed part of the system modelled by EMTDC
to be used by the TS program since TS is only using the fundamental active power
flowing in or out of the interface, the equivalent impedance 2 2 can be selected arbi-
trarily and the current source 7, can be varied to provide the correct power flow.
Moreover, to avoid numerical instability, a constant value of z2, estimated from the
initial power flow results, is used for the duration of the simulation.
8.3 TS5MTDC INTERFACE 301

-
Yes

No
Pass information
to detailed model

i Solve EMTDC

Extract information
from detailed solution
and include in stability
programme

1
1 I
Solve stability equations

c=
T = T + step length

Output results

Is T = end time?

Figure 8.4 Modified TS steering routine

The EMTDC programme represents system 1 by a Thevenin equivalent (El and zl)
as shown in Figure 8.6. The simplest ZI is an R-L series impedance, representing the
fundamental frequency equivalent of system 1. This impedance can be derived from
the results of a power flow and a fault analysis at the interface bus.
The power flow 'provides an initial current through the interface bus and the initial
interface bus. A fault analysis is also used to determine the fault current through the
302 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

f
Figure 8.5 Hybrid interface

I
+

System 1 I,
I
System 2

Figure 8.6 Equivalent circuits at the interface

interface for a short circuit fault to ground. The equivalent Thevenin source El and
Thevenin impedance ZI values shown in Figure 8.7, are obtained as follows:
Using the power flow solution:
El = inZl + V ,
and using the fault circuit: -
El = I F Z ~ .
Combining Equations (8.1) and (8.2):

and can then be found from either Equation (8.1) or Equation (8.2).
During a transient, the impedance of the synchronous machines in system 1 will
change. The net effect on the fundamental power in or out of the equivalent circuit,
however, can be represented by varying the source and keeping ZI constant.
8.3 TS/EMTDC INTERFACE 303

-
(a) Loadflow circuit

- -
(b) Fault circuit

Figure 8.7 Derivation of Thevenin equivalent circuit (a) Power-flow circuit. (b) Fault circuit

However, the waveforms obtained from EMTDC involve frequencies other than the
fundamental and, therefore, a frequency-dependent equivalent circuit (as described in
section 6.4.3)is used to represent the a.c. system.
Regarding the equivalent circuits sources, information from the EMTDC model
representing system 2 (in Figure 8.6) is used to modify the source of the equivalent
circuit of system 2 in the stability programme. Similarly, data from TS are used to
modify the source of the equivalent circuit of system 1 in EMTDC. These equiva-
lent sources are normally updated at each TS step length (refer to section 8.5). From
Figure 8.6,if both ZI and 2 2 are known, additional information is still necessary to
determine update values for the sources 7, and El. The selection and processing of
interface parameters to derive this information is discussed in section 8.4.
Moreover, the stability programme requires only positive sequence data so data from
the three a.c. phases at the interface(s) is analysed and converted to positive sequence,
i.e. for the positive sequence voltage:
-
v,, = p,+ ZVb + Z*V,), (8.4)
where
-
V,, = positive sequence voltage,
- - -
V,, V b , V, = phase voltage,
Z = 120"forward rotation vector(i.e.a = 1L 120").
304 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

Similarly, the positive sequence voltage from TS is converted to three phase data as
follows:
- -
Va = Vps, (8.5)

Some transient stability programmes have the additional capability of modelling


negative and zero sequence networks. If this option is utilized, sequence data from
a stability programme can then be converted to three phase through the following
formulae:
- -
Va = Vps + Vns + Vzs, (8.8)
- -
v b = a 2 v p s + z v n s + v,,, (8.9)
-
v, = zvps+ ans
+ V,,, (8.10)
where
-
Vns = negative sequence voltage,
-
V, = zero sequence voltage.
Similarly, any unbalance in system 2 can be accommodated in the transient stability
programme.

8.3.2 Interface variables derivation


As already explained, in Figure 8.6, El and ZI represent the equivalent circuit of system
1 modelled in EMTDC, while 2 2 and 7, represent the equivalent circuit of system 2
v
modelled in the stability program. is the interface voltage and 71 the current through
the interface which is assumed in the direction shown.
The magnitudes of the interface voltage and current, along with the phase angle
between them, are derived from the detailed EMTDC simulation. This information is
used to update the equivalent circuit source (TC) of system 2 in TS, i.e.:
From Figure 8.6:
-
El = TiZ1 +V , (8.11)
-
v = i2Z2, (8.12)
-
i 2 = i, + 7,. (8.13)
From Equations (8.12) and (8.13):
v = -z1z2
- -
+ icZ2. (8.14)
From Equation (8.1 1):
-
E~ = IIzIL(eII + ezl)+ w e V
=I121cos(6Il + & l ) + j l l Z l sin(& + & I ) +Vcos6v+ jVsin6v (8.15)
8.3 TSEMTDC INTERFACE 305

and
611 = ev - 9, (8.16)
where rb is the displacement angle between the voltage and the current.
Thus, Equation (8.15) can be written as
-
EI = IlZl cos(8v + B ) + jIlZ1 sin(& + B ) + VcosBv + j v s i n b
=IIZl(cos8vcosB-sinevsinB)+ vcosev (8.17)
I Ov cos /? + cosev sin /?)+ V sin ev],
+ ~ [ I I Z(sin
where
B = ezI- 9.
If
Ei = E l , + j E l i ,
then equating the real terms:
E l , = (7,Zl cosB+ V)cos6v + (-1121sinB)sin&, (8.18)
where zlis known and constant throughout the simulation.
From the EMTDC results, the values of V , I and 4 are also known and, hence, so
is j3. El can be determined in the TS phase reference frame from knowledge of zI
and the previous values of interface current and voltage from TS, through the use of
Equation (8.1 1).
From Equation (8.18), making
A = IlZi C O S ~ + V, (8.19)
B = - I I Z ~sin B, (8.20)
and remembering that

A cos ev + B sin ev = J F T 3 c o s ( e v *), (8.21)


where

the voltage angle Bv in the TS phase reference frame can be calculated, i.e.:

(+= tan-' )I:[ . (8.23)

The equivalent current source 7, can be calculated by rearranging Equation (8.14):


- v
I, = -L(ev - eZ2) - Lerl, (8.24)
22
where 811 is obtained from Equation (8.16).
306 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

In a similar way, data from the transient stability programme simulation can be
used to calculate a new Thevenin source voltage magnitude for the equivalent circuit
of system 1 in the EMTDC programme. Knowing the voltage and current magnitude
at the TS programme interface and the phase difference between them, by a similar
analysis the voltage angle in the EMTDC phase reference frame is:

where I,, is the real part of 7, and


V
c = -cosez2 -I, COS~, (8.26)
22
V
D = - cosO~2- I ' sin@, (8.27)
22
4 = ev -ell, (8.28)

+ = tan-' I$[ . (8.29)

Knowing the EMTDC voltage angle 8v allows calculation of the EMTDC current
angle from Equation (8.28). The magnitude value of El can then be derived from
Equation (8.1 1).

8.4 EMTDC to TS Data Transfer


TS assessment is based purely on sinusoidal waveforms, but during faults the actual
waveforms are very distorted. However, there is no consensus on a universal definition
for power when waveforms are non-sinusoidal [ 5 ] .
In the original edition of the book, the total r.m.s. real power was used as the
interfacing variable to transfer information. The total r.m.s. power is made up of
fundamental power (Ff)and harmonic power (&), i.e.:
prms = pf + ph. (8.30)
The r.m.s. power can be extracted directly from the waveforms and, therefore,
Fourier transform or curve fitting methods are not necessary, i.e.:

(8.31)

This method greatly reduces the computing time necessary for the data extraction
from EMTDC.
The choice of r.m.s. power was made on the basis that harmonic power flow will
result only from in-phase components of harmonic voltage and current. The assumption
was made that if a system contains only a low resistive component, then the harmonic
power flow is not significant [6].
This, however, is not valid for every situation and, particularly, at the inverter end of
an HVDC link, the effect of the resistive component of the network is not insignificant.
8.4 EMTDC TO TS DATA TRANSFER 307

Certain harmonic frequencies in a network may also be parallel resonant or close


to parallel resonance and exhibit more resistance than reactance. The presence of a
transient may excite the resonant frequency and greatly affect the results. It is important,
therefore, to model the fundamental power flow accurately.
Another factor to consider is that the direction of fundamental power may not neces-
sarily be the same as the direction of harmonic power. With an HVDC link, while
fundamental power is drawn into a rectifier, much of the harmonic power flow will be
in the reverse direction. The r.m.s. real power measured is the difference between these
two powers and is, therefore, not entirely representative of the fundamental power load
of the HVDC link. Conversely, at the inverter end the r.m.s. real power will include
harmonic power flow into the a.c. system. This may exaggerate the amount of true
fundamental power from the inverter.
As an illustration of the problem, Figure 8.8 shows an EMTDC simulated result of
a single phase fault at the inverter end of the CIGRE HVDC benchmark model [7]
The instantaneous power is shown, along with the total r.m.s. power over discrete one
cycle intervals. For hybrid interfacing purposes, data is transferred at discrete intervals
equal to the stability programme time-step and these are significantly larger than the
interval between the discrete output points constituting the instantaneous power. The
r.m.s. power is then taken over a fundamental period to represent its use in the hybrid
model.
Figure 8.9 shows an analysis of the single-phase fault comparing fundamental
frequency power with the total r.m.s. power. The fundamental frequency power
was derived using the curve fitting method described in section 8.5 to extract both
fundamental voltage and current. The comparison shows that, particularly during the

Instantaneous power -Rms power over T


lo00

800

600

400
F
3 200

[ 0
1
a -200
-400

-600

-800

-1000 , I I I

.O
308 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

1000
900 -
-
800

- 700 -
$ 600-
Y

zi 500-
-m 400-
$ 300-
200 -
100 -
0-
-100 -
-200 I 1

0.00 0.04 0.08 0.12 0.16 0.20 0.24 0.28 0.32 0.36 0.40

Figure 8.9 Fundamental power versus total r.m.s. power

fault time, there exists a significant amount of Ph or harmonic r.m.s. power in the total
r.m.s. power. When the three phase network is unbalanced the fundamental frequency
real power consists of positive, negative and zero sequence components, i.e.:
Pf = Pfps + Pfns + Pfzs, (8.32)
and the negative and zero sequence powers cause additional power loss in the network.
Figure 8.10 shows the sequence components of the fundamental r.m.s. power of
Figure 8.9.
Fundamental frequency negative sequence currents, in the presence of damper wind-
ings, can produce a braking torque which will retard the rotor [8]. Damper windings,
however, also serve to lower the negative sequence impedance of a machine which in
turn reduces the negative sequence voltage [9]. Which of these two opposing effects is
dominant depends on the resistance of the damper windings. High resistance windings
cause the braking torque to be the significant effect. The braking power of the negative
sequence current is:
(8.33)

where
I, = negative sequence rotor current,
R, = rotor resistance,
I,, = negative sequence stator (or armature) current,
8.4 EMTDC TO TS DATA TRANSFER 309

1000 -.
900 -
800 -
F
z 700
600 -
-
' 500 - -
8I 400 -
a 300 -
200 - -
100 - --d
r----

0 - ------ ---- I
..,...r
- - -
-
- - - -

-
-100
-200 - I

Figure 8.10 Sequence components of fundamental frequency r.m.s. power

rn = negative sequence resistance of the machine,


rp = positive sequence resistance of the machine.

The negative sequence resistance can be approximated from the rotor and the arma-
ture resistance, i.e.:
rn % ( ~ s +i ~ r ) ,
(8.35)
where
R, = armature resistance.

Retardation of the rotor can also be caused by d.c. components in the armature
windings. Three phase faults at or near machine terminals can cause d.c. components
of short circuit armature current which can have a definite braking effect on the machine
[9].The braking power in this case is:
(8.36)
(8.37)
where

I , = the effective very low frequency value of the armature current,


idc = instantaneous d.c. component of armature current.

The total r.m.s. power then is not always equivalent toeither the fundamental frequency
power or the fundamental frequency positive sequence power. A comparison of these
310 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

1000
900 -
800 -
-
- 700
600-
v
5 500-
3
-gm 400-
2 300-
200 - Ic..*

100 -
0-
-100- .--
_I.
7-
..

d
-200 I I t I I

0.00 0.04 0.08 0.12 0.16 0.20 0.24 0.28 0.32 0.36 0. 0

Figure 8.11 Comparison of total r.m.s. power, fundamental frequency power and fundamental
frequency positive sequence power

three powers is shown in Figure 8.11. The difference between total r.m.s. power and the
positive sequence power can be seen to be highly significant during the fault.
The most appropriate power to transfer from EMTDC to TS is then the funda-
mental frequency positive sequence power. This, however, requires knowledge of both
fundamental frequency positive sequence voltage and fundamental frequency positive
sequence current. These two variables contain all the relevant information and, hence,
the use of any other power variable to transfer information becomes unnecessary.

8.5 Data Extraction from Distorted Waveforms


At each step of the transient stability program power transfer information needs to be
derived from the distorted current and voltage waveforms at the points of interface.
This can be achieved using the FFT which provides accurate information for the whole
frequency spectrum. However, only the fundamental frequency is used in the stability
program and a simpler curve fitting approach is described next that provides sufficient
accuracy.
A curve fitting algorithm (CFA) can be used to extract the fundamental frequency
data, based on a least squared error technique. It can be described as follows: Assume
a sinewave signal with a frequency of o radians per second and a phase shift of $
relative to some arbitrary time To.
y ( t ) = A sin(@?- $), (8.38)
where +=d o .
8.5 DATA EXTRACTION FROM DISTORTED WAVEFORMS 311

This can be rewritten as:


y ( t ) = A sin(wt) cos(wT0) - A cos(ot) sin(wT0). (8.39)
Letting C1 = A cos(wTo) and C2 = A sin(wT0) and if sin(wt) and cos(ot) are repre-
sented by functions F I ( t ) and F2(t), respectively, then:
y(t> = CI FI ( t ) + C2F2(t). (8.40)
F l ( t ) and F 2 ( t ) are known if the fundamental frequency w is known. However, the
amplitude and phase of this frequency generally need to be found, so the equation has
to be solved for C I and C2. If the signal y ( t ) is distorted, then its deviation from a
sinusoid can be described by an error function E :
~ ( t=) y ( t ) +E . (8.41)
For a least squares method of curve fitting, {he size of the error function is measured
by the sum of the individual residual squared values such that:

i= 1
+
where x; = x(t0 i A t ) and yi = y(t0 + iAr).
From Equation (8.40):
n
E= C ( x ; - ~ I ~ l ( t-i ~) 2 ~ 2 ( t i ) ) ~ , (8.43)
i= I

where the residual value r at each discrete step is defined as:

In matrix form:

(8.45)

or

The error component can be described in terms of the residual matrix as follows:
E = [rlT[r]
312 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

This error then needs to be minimized.

(8.48)
If [A] = [FIT[F] and [B] = [FIT[X] then :
[CI = [ A I - m (8.49)

and hence

Elements of matrix [A] can then be derived as shown:

n-1

(8.50)

Similarly:

and

(8.5 1)
(8.52)

From these matrix element equations, CIand C2 can be calculated recursively using
sequential data.
8.6 INTERFACE METHOD 313

2oo I

-5.00 I 1 I I I 1 I 1 I 1

0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20
Time (s)

Figure 8.12 Simulated fault waveform

8.5.1 CFA effectiveness


The accuracy of the CFA algorithm has been compared with the results of an FFT
for the voltage waveform of Figure 8.12. This waveform, derived with the EMTDC
program, corresponds to a single-phase fault on the rectifier side of the CIGRE bench-
mark model [7].
Several windows were used to capture the fundamental frequency information. Of
these, a discrete rectangular window of full period duration provides the best match
when data are only required at intervals of one cycle. However, if data are required at
more frequent intervals, a staggered (or overlapping) window producing results twice
per cycle is a better choice.

8.6 Interface Method


The data from each program must be interchanged at appropriate points during the
hybrid simulation run. The timing of this data interchange between the TS and EMTDC
programs is important, particularly around discontinuities caused by fault application
and removal.
The electromechanical a.c. system solution exhibits a relatively slow dynamic
response to any disturbances and the accuracy of a previous time-step information
passed from TS to EMTDC is, therefore, an adequate assumption. On the other hand,
the HVDC converter exhibits fast dynamic behaviour. Present time-step information
is, thus, more appropriate to pass from EMTDC to TS when solving the TS network
equation. Therefore, of the different possible interfacing alternatives, the method
314 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

illustrated in Figure 8.13 is selected for the information exchange. It shows the use
of the previous time-step data when passing information from TS to EMTDC but
present time-step data when passing information from EMTDC to TS. When the two
programs are run concurrently, TS passes its information to EMTDC. EMTDC then
runs to t + A t , and the information gathered over this present time-step given back to
+
TS at time t . TS now runs to t Ar, using equivalent circuit information derived for
that particular time step.
However, to cater for TS step lengths below the fundamental period, the above
method is modified as shown in Figure 8.14; it relates to the case when the step length
is one half of the fundamental period.
Following the sequential numbering on Figure 8.14, at a particular point in time,
the EMTDC and TS programs are concurrent and the TS information from system 1
is passed to update the system 1 equivalent in EMTDC. This is shown by the arrow
marked 1. EMTDC is then called for a length of half a fundamental period (arrow
2) and the curve fitted results over the last full fundamental period processed and
passed back to update the system 2 equivalent in TS (arrow 3). The information over
this period is passed back to TS at the mid-point of the EMTDC analysis window
which is half a period behind the current EMTDC time. TS is then run to catch up

I I

Figure 8.13 Interfacing methods

Stabiliity

--
step length

I I
4
I
I
8
I
I
etc.
I
1-1 I I
I TS

-- 2 6 4 c III I
Dynamic program
step length

Figure 8.14 Normal interaction protocol


8.7 INTERFACE LOCATION 315

Disturance Stabiliity
applied step length

I
I
l
I
i
I
4
l
I
t
I
*
I
I
I
- /-I
I
8 etc.
I I I
TS

v p * t
++ II
II
r Dynamic program
step length

Figure 8.15 Interaction protocol around a disturbance

to EMTDC (arrow 4), and the new information over this simulation run used to again
update the system 1 equivalent in EMTDC (arrow 5 ) . This protocol continues until
any discontinuity in the network occurs.
When a network change such as a fault application or removal occurs, the interaction
protocol is modified to that shown in Figure 8.15. The curve fitting analysis process is
also modified to avoid applying an analysis window over any point of discontinuity.
The sequential numbering in Figure 8.15 explains the flow of events. At the fault
time, the interface variables are passed from TS to the system 1 equivalent in EMTDC
in the usual manner, as shown by the arrow marked 1. Neither system 1 nor system
2 has yet been solved with the network change. The fault is now applied in EMTDC
which is then run for a full fundamental period length past the fault application (arrow
2) and the information obtained over this period passed back to TS (arrow 3). The fault
is now also applied to the TS program which is then solved for a period until it has
again reached EMTDCs position in time (arrow 4). The normal interaction protocol
is then followed until any other discontinuity is reached.
A full period analysis after the fault is applied is necessary to accurately extract
the fundamental frequency component of the interface variables. The mechanically
controlled nature of the a.c. system implies a dynamically slow response to any distur-
bance and so, for this reason, it is considered acceptable to run EMTDC for a full
period without updating the system 1 equivalent circuit during this time.

8.7 Interface Location


The original intention of the initial hybrid algorithm [6] was to model the a.c. and d.c.
solutions separately. The point of interface location was consequently the converter
bus terminal. The detailed d.c. link model included all equipment connected to the
converter bus, such as the a.c. filters, and every other a.c. component was modelled
within the stability analysis. A fundamental frequency Thevenin equivalent was used
to represent the stability program in the detailed solution and vice versa.
316 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

An alternative approach has been proposed [ 101 where the interface location is
extended out from the converter bus into the a.c. system. This approach maintains that,
particularly for weak a.c. systems, a fundamental frequency equivalent representing the
a s . system is not adequate at the converter terminals. In this case, the extent of the a.c.
system to be included in the d.c. system depends on phase imbalance and waveform
distortion.
Although the above concept has some advantages, it also suffers from many disad-
vantages. The concept is proposed, in particular, for weak a.c. systems. A weak a.c.
system, however, is likely to have any major generation capability far removed from the
converter terminal bus as local generation serves to enhance system strength. However,
when the generation is far away from the converter, then the distance required for an
interface location to achieve considerably less phase imbalance and waveform distortion
is also likely to be significant.
The primary advantage of a hybrid solution is in accurately providing the d.c.
dynamic response to a transient stability program, and in efficiently representing the
dynamic response of a considerably sized a.c. system to the d.c. solution. Extending the
interface some distance into the a.c. system, where the effects of a system disturbance
are almost negligible, diminishes the hybrid advantage. If a sizeable portion of the a.c.
system requires modelling in detail before an interface to a transient stability program
can occur, then one might question the use of a hybrid solution at all and instead use
a more conventional approach of a detailed solution with a.c. equivalent circuits at the
system cut-off points.
Another significant disadvantage in an extended interface is that ax. systems may
well be heavily interconnected. The further into the system that an interface is
moved, the greater the number of interface locations required. The hybrid interfacing
complexity is thus increased and the computational efficiency of the hybrid solution
decreased. The requirement for a detailed representation of a significant portion of
the a.c. system serves to decrease this efficiency, as does the increased amount of
processing required for variable extraction at each interface location.
The advantages of using the converter bus are:
0 The detailed system is kept to a minimum.
0 Interfacing complexity is low.
0 Computational expense is minimized.
0 Converter terminal equipment, such as filters, synchronous condensers SVCs, and
can still be modelled in detail.
The major drawback of the detailed solution is in not seeing a true picture of the
ax. system, since the equivalent circuit is fundamental frequency based. Waveform
distortion and imbalance also make it difficult to extract the fundamental frequency
information necessary to transfer to the stability program.
The problem of waveform distortion for transfer of data from EMTDC to TS is
dependent on the accuracy of the technique for extraction of interfacing variable infor-
mation. If fundamental frequency quantities can be accurately measured under distorted
conditions, then the problem is solved. Section 8.5 has described an efficient curve
fitting algorithm to extract the required information from distorted waveforms. It has
8.8 STRUCTURE OF THE HYBRID PROGRAM 317

Rectifier 7 1 Inverter ESCR=2.0 h+-@

Figure 8.16 Test system

been shown that, using the technique described, fundamental frequency quantities can
be accurately measured.
Moreover, a simple fundamental frequency equivalent circuit is insufficient to repre-
sent the correct impedance of the a.c. system at each frequency. Instead, this can be
achieved by using a fully frequency-dependent equivalent circuit of the a.c. system
[ l 11 at the converter terminal instead of just a fundamental frequency equivalent. A
frequency-dependent equivalent avoids the need for modelling any significant portion
of the ax. system in detail yet still provides an accurate picture of the system impedance
across its frequency spectra.
To show the effect of a frequency-dependent equivalent, the test system of
Figure 8.16 is used with an inverter effective short circuit ratio of 2. The d.c. link
is represented by the CIGRE benchmark model [7].
A three-phase solid short circuit fault is applied at the inverter terminal. Three cases
are considered, the first being the entire system represented by the detailed solution.
The second and third cases relate to the hybrid solution interfaced at the converter
bus, one case with a fundamental frequency Thevenin representation of the stability
program in the detailed solution, and the other with a frequency-dependent equivalent.
The results of the inverter terminal voltages for the three cases are plotted in
Figure 8.17, and they show that the benchmark EMTDC case and the frequency-
dependent equivalent case are identical. The fundamental frequency equivalent case
(Figure 8.17(b)) shows more distortion and prolonged effects from the disturbance
than the benchmark EMTDC case.
These results show the inadequacy of a fundamental frequency equivalent at the
converter terminal, whereas the use of a frequency-dependent equivalent is perfectly
adequate.

8.8 Structure of the Hybrid Program


The hybrid algorithm interface is shown in Figure 8.18. It contains the five possible
states summarized in Table 8.1. The default state is 0 for the first time that the algorithm
is called. This occurs once TS has reached steady state electromechanical equilibrium
and is ready for interfacing, prior to a system disturbance.
Under the 0 state, EMTDC is called for a full fundamental period ahead of TS and
its interfacing variables extracted and converted to positive sequence. The state is now
set to -1.
Under the - 1 state, TS is run for one half of a fundamental period and the variables
at its interface location are compared with those of EMTDC. This is simply to check
and ensure that the data in both files is correctly set up. The TS network is then split
318 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

-1.4 I i i i i i i i i

(a) EMTDC

1.4 , i

1.2
1 .o
0.8

-& 0.6
0.4
0.2
g) 0.0
-0.2
>o -0.4
-0.6
-0.8
-1 .o
-1.2
-1.4 I 1 , , 1 1 1 1 1

(b) TSE Hybrid-Fundamental frequency equivalent

1.4 , f

0.96 1.00 1.02 1.04 1.06 1.08 1.10 1.12 1.14 1.16 1.18
Time (s)
-
(c) TSE Hybrid Frequency dependent equivalent

Figure 8.17 Inverter ax. voltage


8.8 STRUCTURE OF THE HYBRID PROGRAM 319

Table 8.1 Interfacing states


State Event
-1 Isolate TS network ready to interface with EMTDC
0 Initial time through interfacing routine
1 Normal interaction between TS and EMTDC
2 System network change signalled-run EMTDC for a full period
past the fault and return information to TS
3 TS catching up with EMTDC after a system network change

No
1
Yes 1 1
Compare variables I
s = Ts+stepl
I
I

No 1
1 and2

Set up equivalent
circuits

Yes State=l
Ts=O

Call EMTDC for a

Convert EMTDC olp

Ts=Ts+steol

A 0

Figure 8.18 TSE hybrid algorithm. [step 11 TS program step length. [Ts] TS step length
counter. [Tend] Minimum finish time for EMTDC. [K] Time counter for EMTDC and TS
comparison of variables. [Tc] Time over which comparison must be within tolerance
320 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

Yes I

eriod of TI2

variables to TS sys. 1

State=3
Ts=O
Ts=Ts+stepl
~~~~~~~~ ~

Extract TS system
2 variables
lr
Compare variables of
TS sys.2 and EMTDC

+=+&

I- K=K+stepl
No
i

Terminate EMTDC
*
Reconnect TS
systems 1 and 2
I I

C D

Figure 8.18 (Continued)


8.8 STRUCTURE OF THE HYBRID PROGRAM 321

IC
stop

1 f
Pass TS,System 1 1 f
variables to EMTDC
% Ts=Ts+step 1

Ts=O

Figure 8.18 (Continued)

into systems 1 and 2, and Norton equivalent circuits set up in each system. The state
is now set to 1, representing normal interfacing conditions, and TS is solved until its
time is concurrent with EMTDC.
Under the normal interfacing of state 1, variable information is given to EMTDC
from TS to update its equivalent circuit. The equivalent circuit of system 1 in the TS
model of system 2 is updated in the same way. EMTDC is then called for one half a
period and the information measured over the last fundamental period of its simulation
. . . . . . , -
used to update the equivalent circuit or system L in 15.
a . rn"

In the interface data file, a time can be specified as a minimum termination time
for EMTDC, after any network disturbance is cleared. Once this time is reached in
the program, the interface variables of both system 2 models are compared at each
TS time-step. If they correlate within a predefined tolerance over a set period, then
EMTDC can be terminated and the TS representation of system 2 once again takes
river nurino
" 1 1 1 . ..." nerinrl
b the
Y"1
1.. y, .
..,.. thpv
-.."J rirp hpino
I.w" -..--.-..-,if_- at anv,time
-... rhprkpd I---f --.-.-
the
_-.-vsriahleq
.
-_---_-I -_- outside
are -
the specified tolerance, the checking period time is restarted.
322 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

When a network admittance change is about to occur, such as a fault being applied
or removed, the interfacing routine is entered with a state value of 2. At this point
in time, TS is concurrent with EMTDC. The TS interface variables are passed to
EMTDC, which is then run for a full period from the network discontinuity time. It
is not accurate to apply the analysis window over the network discontinuity, so the
staggered window process must be restarted. The variables from EMTDC are passed
back to TS at the time of the disturbance and the state is changed to 3.
State 3 allows TS to catch up to EMTDC before the normal interfacing procedure
of state 1 recommences.

8.9 Test System and Results


The d.c. link of the test system is based on the CIGRE benchmark model [7] described
in Appendix 111. However, the rectifier a.c. system is modified as shown in Figure 8.19
to be representative of the New Zealand central South Island primary transmission
system. The parameters for this test system are also given in Appendix 111. A three-
phase short circuit is applied to the rectifier terminals of the d.c. link (at time 1.7 s) and
cleared three cycles later. For the EMTDC studies, the a.c. system at the faulted (recti-
fier) end is replaced by a frequency-dependent equivalent circuit, derived as explained
in section 6.4.3, which produced the harmonic impedances, shown in Figure 8.20.

8.9.1 Response of the individual programs


The differences between the TS and EMTDC programs when used independently from
each other are illustrated in Figure 8.21
Following fault clearance the rectifier is unblocked at t = 1.8s and a few ms later
the inverter takes over conduction from its by-pass valves. After an initial current
inrush into the d.c. line capacitance, the power setting is ramped up. At t = 1.97 s the
inverter is restated and the link current drops, a reduction which is more pronounced
and stays longer in the EMTDC solution.

Clyde Twizel

Livingston Aviemore

Figure 8.19 Test system on the a.c. side of the rectifier station
8.9 TEST SYSTEM AND RESULTS 323

2 3000
-3 2000-
r
o 2500- t
-6
.-
1500-
ii

m
E 1000-
8 500 -
2
0

1.20
7 1.00
n
0.80-
C
0.60-
3
0.40-
0
0 0.20-
0.00
.....
I I I I I I I I
1.50 1.65 1.80 1.95 2.10 2.25 2.40 2.55 2.70 2.85 3.00
Time (s)

Figure 8.21 d.c. current response to a three-phase short circuit at the rectifier end

Following the power ramp, the a.c. current returns to its nominal value and after a
small oscillatory settling of the controllers so do the d.c. voltage and current.

8.9.2 TSE hybrid response


The results of the EMTDC and TS simulations show substantial differences and thus
their independent use is inadequate to model the transient stability of the a.c. system.
Figure 8.22(a) displays the rectifier terminal a.c. voltage levels calculated with the
TS and TSE alternatives. The main difference between them is a transient overshoot
in the TSE solution immediately after fault clearance.
The rectifier d.c. currents, displayed for the three solutions in Figure 8.22(b), show
a very similar variation for the TSE and EMTDC solutions, except for the region
between I = 2.03 s and t = 2.14 s but the difference with the TS only solution is very
pronounced.
Figure 8.23 compares the fundamental positive sequence real and reactive powers
across the converter interface for the TS and TSE solutions.
The main differences in real power occur during the link power ramp. The difference
is almost a direct relation to the d.c. current difference between TS and TSE shown
in Figure 8.22(b). The oscillation in d.c. voltage and current as the rectifier terminal
is de-blocked is also evident.
324 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

................ TS only T S E (TS variable)


1.40 I 1
1.20
1 .oo
0.80
0.60
0.40
0.20
0.00

1.20
1.20 -- _I

z 1.00 -
g 0.80-
6 0.60-
0 0.40 -
0.20 -
0.00 I I I I I I I I
1.50 1.65 1.80 1.95 2.10 2.25 2.40 2.55 2.70 2.85 3.00
(b) Time (s)

Figure 8.22 Hybrid response comparisons. (a) Rectifier end a x . voltage. (b) Rectifier end d.c.
current

- TSE TS
1200 , I
1000

58 600
400

0
I I I I I I I I I

d
h
200

0
-200-
-- .......... ...............
~

z. -400-
p8 -600-

.-rg -800-
5 -1Ooo-
I
a -1200 -
-1400 I I I I I I I I I

Figure 8.23 Real and reactive power across interface


8.10 QUASI STEADY-STATE CONVERTER SIMULATION 325

- -
5
50
30 --
- - - - - --
I ,
I ~.,~,-I~,.R-,,-~--.
1,
I -- -_-.- - - - - - -
s -
-a lo-
pi F,
.,i \...!. !..,/*\,.,..-.
0,
5 -10-
........... ......................................................................
L........
a
c -30-
E
0
-50

Figure 8.24 Rotor angle swings using TSE

As for the reactive power Q, prior to the fault, a small amount is flowing into
the system due to a surplus MVArs at the converter terminal. The fault reduces this
power flow to zero. When the fault is removed and the a.c. voltage overshoots in TSE,
the reactive MVArs also overshoot in TSE and since the d.c. link is shut down, a
considerable amount of reactive power flows into the system.
Finally, Figure 8.24 shows the machine angle swings with respect to the Clyde
generator (see test system of Figure 8.19). These indicate that the system is transiently
stable.

8.10 Quasi Steady-state Converter Simulation


When analysing small perturbations and dynamic stability, converter equipment is
expected to operate in a controlled manner almost instantaneously when compared
with the relatively slow a.c. system dynamics. In these cases, it is acceptable to use a
modified steady-state (or quasi steady-state) model, the modifications being due to the
different constraints imposed by the load flow and stability studies. Such a model is
also suitable for representing large rectifier loads during a.c. system disturbances, with
further modifications necessary to represent abnormal rectifier operating modes.
Further to the basic assumptions listed in Chapter 3, the following need to be made
here:
0 The implementation of delay angle control is instantaneous.
0 The transformer tap position remains unchanged throughout the stability study
unless otherwise specified.
0 The direct current is smooth, though its actual value may change during the study.

8.10.1 Rectifier loads


Large rectifier loads generally consist of a number of series andor parallel connected
bridges, each bridge being phase shifted relative to the others. With these configu-
rations, high-pulse numbers can be achieved resulting in minimal distortion of the
326 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

Figure 8.25 Rectifier load equivalent circuit

Reducing aV,,,,cosa

Constant current
control characteristic

;
.
'd* 'd

Figure 8.26 Simple rectifier control characteristic

supply voltage without filtering. Rectifier loads can, therefore, be modelled as a single
equivalent bridge with a sinusoidal supply voltage at the terminals but without repre-
sentation of passive filters. This model is shown in Figure 8.25.
Rectifier loads can utilize a number of control methods which can be modelled using
a controlled rectifier with suitable limits imposed on the delay angle (a)[12].
Static loads Operating under constant current control, the d.c. equations are:
(8.53)

(8.54)

where A is the constant current controller gain and I d s is the normal d.c. current setting
as shown in Figure 8.26.
Constant current cannot be maintained during a large disturbance as a limit of delay
angle will be reached. In this event, the rectifier control specification will become one
8.10 QUASI STEADY-STATE CONVERTER SIMULATION 327

of constant delay angle and Equation (8.54) becomes:

(8.55)

Protection limits and disturbance severity determine the rectifier operating charac-
teristics during the disturbance. Shutdown occurs if I d reaches a set minimum or zero
and the voltage Vload will cause shutdown before the a.c. terminal voltage reaches zero.
The action of the rectifier load system is thus described by Equations (3.24), (3.30),
(3.34), (3.36), (3.37), (8.53) and either (8.54) or (8.55).
Dynamic loads The basic rectifier load model assumes that current on the d.c. side
of the bridge can change instantaneously. For some types of rectifier loads this may
be a valid assumption, but the d.c. load may well have an overall time constant which
is significant with respect to the fault clearing time. In order to examine realistically
the effects which rectifiers have on the transient stability of the system, this time
constant must be taken into account. This requires a more complex model to account
for extended overlap angles, when low commutating voltages are associated with large
d.c. currents.
When the delay angle (a)reaches a limiting value, the dynamic response of the d.c.
current is given by
V d = IdRd + +
Vload Ldpld, (8.56)

where 4 represents the equivalent inductance in the load circuit. Substituting for Vd
using Equation (3.24) gives

where T d c = L d / R d .
The controller time constant may also be large enough to be considered. However,
in transient stability studies where large disturbances are usually being investigated,
faults close to the rectifier load force the delay angle (a)to minimum very quickly.
Provided the rectifier load continues to operate, the delay angle will remain at its
minimum setting throughout the fault period and well into the post-fault period until
the terminal voltage recovers. The controller will, therefore, not exert any significant
control over the d.c. load current. Ignoring the controller time constant can, therefore,
be justified in most studies.

Abnormal modes of converter operation The slow response of the d.c. current
when a large disturbance has been applied to the ax. system can cause the rectifier to
operate in an abnormal mode.
After a fault application near the rectifier, the near normal value of d.c. current
( I d ) needs to be commutated by a reduced a.c. voltage. This causes the commutation
angle ( b )to increase and it is possible for it to exceed 60". This mode of operation is
beyond the validity of the equations and to model the dynamic load effects accurately
it is necessary to extend the model.
328 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

The full range of rectifier operation can be classified into four modes [13]:
Mode 1 -Normal operation. Only two valves in the bridge are involved in simulta-
neous commutation at any one time. This mode extends up to a commutation angle
of 60".
Mode 2 -Enforced delay. Although a commutation angle greater than 60" is desired,
the forward voltage across the incoming thyristor is negative until either the previous
commutation is complete or the firing angle exceeds 30". In this mode, p remains
at 60" and a ranges up to 30".
Mode 3 -Abnormal operation. In this mode, periods of three-phase short circuit and
d.c. short circuit exist when two commutations overlap. During this period, there
is a controlled safe short circuit which is cleared when one of the commutations is
complete. During the short-circuit periods, four valves are conducting. Commutation
cannot commence until 30" after the voltage crossover.
Mode 4 -Continuous three-phase and d.c. short circuit caused by two commutations
taking place continuously. In this mode, the commutation angle is 120" and the a.c.
and d.c. current paths are independent.

The waveforms for these modes are shown in Figure 8.27, and Table 8.2 summarizes
the conditions for the different modes of operation. Equations (3.24) and (3.30) do not
apply for a rectifier operating in Mode 3 and they must be replaced by:

(8.58)

and
a
Id = -Vterm(c0sa'- cos y'). (8.59)
45xc
Fourier analysis of the waveform leads to the relationship between ax. and d.c.
current given by Equation (3.32), where the factor k is now:

(8.60)

where
a' = CY - 30"
and
y' = y + 30". (8.61)
A graph showing the value of k for various delay angles (a)and commutation angles
( p ) is shown in Figure 8.28.

Identification of operating mode The mode in which the rectifier is operating can
be determined simply by use of a current factor K I . The current factor is defined as:

(8.62)

Substitution in this, using the relevant equations, yields limits for the modes.
8.10 QUASI STEADY-STATE CONVERTER SIMULATION 329

6 2 4 6

2 4 6

Figure 8.27 Rectifier voltage waveforms showing different modes of operation. (a) Mode 1,
g < 60"; (b) Mode 2, p = 60" with enforced delay al; (c) Mode 3, p z 60" with short-circuit
period a2

Table 8.2 Rectifier modes of operation


Mode Firing angle Overlap angle
1 0" 5 a 5 90" 0" 5 /A 5 60"
2 0" 5 a 5 30" 60"
3 30" 5 (Y 5 90" 60" 5 p < 120"
4 30" 5 a 5 90" 120"
330 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

?
I 1 I 1

1.2 -

Figure 8.28 Variation of k in expression I , = k ( 3 , h / n ) I d

Mode 1:
K I 5 COS(~O"- a ) (8.63)
and
K I 5 2 cos a for rectifier operation.
Mode 2.
(8.64)
L

Mode 3:
2
K I < - when a 5 30,
43
2
K I < -cos(ct - 30") when ct > 30". (8.65)
43
Mode 4:
2
K I = - when a 5 30,
43
2
K I = -cos(a - 30") when a > 30" (8.66)
43
This can be demonstrated in the curve of converter operation shown in Figure 8.29.
It can thus be seen that the mode of operation can be established prior to solving
for the rectifier load equations at every step in the solution.

8.10.2 d.c. link


Provided that it can be safely assumed that a d.c. link is operating in the quasi steady-
state (QSS) Mode 1, the equations developed for converters in Chapter 3 can be used.
That is, the converters are considered to be controllable and fast acting so that the
8.10 QUASI STEADY-STATE CONVERTER SIMULATION 331

Mode 4

Delay angle, a

Figure 8.29 Converter operation [ 141

normal steady-state type of model can be used at each step in the transient stability
study.
The initial steady-state operating conditions of the d.c. link will have been deter-
mined by a load flow and, in this, the control type, setting and margin will have been
established.
During the solution process at each iteration, the control mode must be established.
This can be done by assuming Mode 1 (i.e. with the rectifier on C.C. control) and,
by combining Equations (3.39), (3.45), (8.53) and (8.54), a d.c. current can be deter-
mined as:
Idsr - [(3fi/n)aiVtermi cos Yicl/Ar
I d mode I = (8.67)
+
1 (Rd - (3/n)xci)/Ar
Assuming this current to be valid, then d.c. voltages at each end of the link can be
calculated using Equations (3.24) and (3.39).
The d.c. link is operating in mode 2 (i.e. with the inverter on C.C. control) i f
Vdr mode I - Vdi mode I =< 0. (8.68)
The d.c. current for mode 2 operation is given by

For constant power control, under control mode 1, the d.c. current may be determined
from the quadratic equation:

where Pds is the setting at the electrical mid-point of the d.c. system, i.e.
pds = (Pdsr + pdsi)/2* (8.7 1)
332 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

Table 8.3 Current setting for constant power control


from quadratic equation

Id I Id2

Within Outside Id I
Outside Within Id2
Within Within Greater of I d , and Ir12
Greater Greater I d max
Greater Less I d max
Less Greater I d max
Less Less 0
Within = within the range / d m i n to / d m a x ;
Outside = outside the range l d m i n to l d m a x ;
Greater c Greater than I d m a x ;
Less = Less than I d m i " .

The correct value for Idm&I can then be found from Table 8.3. Control mode 2
is determined using Equation (8.68) and in this case the following quadratic equation
must be solved:
k r z : mode 2 - k v l d mode 2 - P d marg - p d s = 0, (8.72)
where

(8.73)

(8.74)

If the link is operating under constant power control but with a current margin then
for control mode 2:

It is possible for the d.c. link to be operating in control mode 2 despite satisfying the
inequality of Equation (8.68). This occurs when the solution indicates that the rectifier
firing angle (a,)is less than the minimum value ( a r m i n ) . In this case, the delay angle
should be set to its minimum and a solution in mode 2 is obtained.
It is also possible that when the link is operating close to the changeover between
modes, convergence problems will occur in which the control mode changes at each
iteration. This can easily be overcome by retaining mode 2 operation whenever detected
for the remaining iterations in that particular time step.
d.c. power modulation It has been shown in the previous section that under the
constant power control mode, the d.c. link is not responsive to a.c. system terminal
conditions, i.e. the d.c. power transfer can be controlled disregarding the actual a.c.
voltage angles. Since, generally, the stability limit of an ax. line is lower than its
thermal limit, the former can be increased in systems involving d.c. links, by proper
utilization of the fast converter controllability.
The d.c. power can be modulated in response to a.c. system variables to increase
system damping. Optimum performance can be achieved by controlling the d.c. system
8.10 QUASI STEADY-STATE CONVERTER SIMULATION 333

so as to maximize the responses of the a.c. system and d.c. line simultaneously
following the variation of terminal conditions.
The dynamic performance under d.c. power modulation is best modelled in three
separate levels [14]. These levels, illustrated in Figure 8.30, are (a) the a.c. system
controller, (b) the d.c. system controller and (c) the ax.-d.c. network.

(a) The ax. system controller uses ax. andor d.c. system information to derive the
current and voltage modulation signals. A block diagram of the controller and
a.c.-d.c. signal conditioner is shown in Figure 8.31.
(b) The d.c. system controller receives the modulation signals AI and A E and the
steady-state specifications for power PO current I 0 and voltage Eo. Figure 8.32(a)
illustrates the power controller model, which develops the scheduled current
setting; it is also shown that the current order undergoes a gradual increase during
restart, after a temporary blocking of the d.c. link.
The rectifier current controller, Figure 8.32(b), includes signal limits and rate
limits, transducer time constant, bandpass filtering and a voltage dependent current
order limit (VDCOL).
The inverter current controller, Figure 8.32(c), includes similar components plus a
communications delay and the system margin current (Z,,,). Finally, the d.c. voltage
controller, including voltage restart dynamics, is illustrated in Figure 8.32(d).
(c) The d.c. current I d and voltage E d derived in the d.c. system controller constitute
the input signals for the a.c.-d.c. network model which involves the steady state

ffY
DC control modes Ed 'd 'd, ' ' d i
I_ AClDC interface Qr Ql

(a) (b) (4
Figure 8.30 a.c.-d.c. dynamic control structure: (i) a.c. system controller, (ii) d.c. system
controller, (iii) ax.-d.c. network

Figure 8.31 a.c. system controller


334 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

N(P)
3~- '07

D(P)

Figure 8.32 d.c. system controller. (a) Power controller; (b) Rectifier current controller;
(c) Inverter current controller; (d) d.c. voltage controller

solution of the d.c. system (neglecting the d.c. line dynamics which are included
in the d.c. system controller). Here the actual a.c. and d.c. system quantities are
calculated, i.e. control angles, d.c. current, voltage, active and reactive power.
The converter a.c. system constraints are the open circuit secondary voltages Ear
and Eai.

8.10.3 Representation of converters in the network


Rectifiers The static load rectifier model can be included in the overall solution
of the transient stability program in a similar manner to the basic loads described in
Chapter 7.
8.10 QUASI STEADY-STATE CONVERTER SIMULATION 335

From the initial load flow, nominal bus shunt admittance (yo) can be calculated
for the rectifier. This is included directly into the network admittance matrix [Y].The
current injected into the network in the initial steady state is, therefore, zero. In general:

where
?;*
- 30
(8.77)
Yo=Ivrermo12
and
(8.78)

The static load rectifier model does not depart greatly from an impedance charac-
teristic and is well behaved for low terminal voltages, the injected current tending to
zero as the voltage approaches zero. Figure 8.33 compares the current due to a rectifier
with that due to a constant impedance load. As the injected current is never large, the
iterative solution for all ax. conditions is stable.
When the rectifier model is modified to account for the dynamic behaviour of the
d.c. load, its characteristic departs widely from that of an impedance. Immediately after
a fault application, the voltage drops to a low value but the injected current magni-
tude does not change significantly. Similarly, on fault clearing, the voltage recovers
instantaneously to some higher value while the current remains low.

Ip t

I
0.5 1.o
Vterm
linj 4
~F.L. -

Figure 8.33 Difference between impedance and static load rectifier characteristic (for
vload # 0)
336 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

When the load characteristic differs greatly from that of an impedance, a sequential
solution technique can exhibit convergence problems, especially when the voltage is
low. With small terminal voltages, the a.c. current magnitude of the rectifier load is
related to the d.c. current but the current phase is greatly affected by the terminal
voltage. Small voltage changes in the complex plane can result in large variations of
the voltage and current phase angles.
To avoid the convergence problems of the sequential solution, an alternative algo-
rithm has been developed [ 151. This combines the rectifier and network solutions into
a unified process. However, it does not affect the sequential solution of the other
components of the power system with the network.
The basis of this approach is to reduce the a.c. network, excluding the rectifier, to an
equivalent Thevenin source voltage and impedance as viewed from the primary side
of the rectifier transformer terminals. This equivalent of the system, along with the
rectifier, can be described by a set of non-linear simultaneous equations which can be
solved by a standard Newton-Raphson algorithm. The solution of the reduced system
yields the fundamental a.c. current at the rectifier terminals.
To obtain the network equivalent impedance, it is only necessary to inject 1 p.u.
current into the network at the rectifier terminals while all other nodal injected currents
are zero. With an injected current vector of this form, a solution of the nodal network
Equation (7.48) gives the driving point and transfer impedances in the resulting voltage
vector:
[Z] = [V']= [T]-l[7;,1, (8.79)

(8.80)

The equivalent circuit shown in Figure 8.34 can now be applied to find the rectifier
current (7,) by using the Newton-Raphson technique.
The effect of the rectifier on the rest of the system can be determined by superpo-
sition:
+
[VI= [TI [ZII,, (8.81)
where
[TI= [r]-"l;j] (8.82)

and [7Ynj]are the injected currents due to all other generation and loads in the system.
If the network remains constant, vector [z] is also constant and thus only needs
re-evaluation on the occurrence of a discontinuity.
Thus, the advantages of the unified and sequential methods are combined. That is,
good convergence for a difficult element in the system is achieved while the program-
ming for the rest of the system remains simple and storage requirements are kept low.
8.10 QUASI STEADY-STATE CONVERTER SIMULATION 337

'I
Rd

vd=0

Dynamic
load element

2 Vload

Figure 8.34 Equivalent system for Newton-Raphson solution

The equivalent system of Figure 8.34 contains seven variables ( V , e r m , I , , 8, @, a,


Vd and I d ) . With these variables, four independent equations can be formed. They are
Equation (3.24) and:
Vl!@- V t e r m L 8 - Z,hl!(IpL$' = 0, (8.83)
Vdld - &aVtermlp cOS(6 - 9) = 0. (8.84)
Equation (8.83) is complex and represents two equations. Substituting for V d and
I , using Equations (3.34) and (8.53) reduces the number of variables to five. A fifth
equation is necessary and with constant current control, i.e. with the delay angle (a)
within its limits, this can be written as
Id - I d s p = 0. (8.85)
Equation (3.24), suitably reorganized, and Equations (8.83)-(8.85)represent [ F ( X ) ]=
0 of the Newton-Raphson process and
[XIT = [Er, 8, a,I d ] . $9 (8.86)
When the delay angle reaches a specified lower limit (omin),the control specification,
given by Equation (8.85) changes to
a - amin = 0. (8.87)
Equation (3.24) is no longer valid. The d.c. current ( I d ) is now governed by the
differential Equation (8.57). If the trapezoidal method is being used, this equation
can be transformed into an algebraic form similar to that described in Chapter 7 .
Equation (3.24) is replaced by:
Id = kaE, cosa - kb = 0. (8.88)
The variables ka and kb contain information from the beginning of the integration
step only and are thus constant during the iterative procedure:
ka = h / ( 2 + kc h), (8.89)
338 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

where
(8.91)

t represents the time at the beginning of the integration step and h is the step length.
Commutation angle p is not explicitly included in the formulation, and since these
equations are for normal operation, the value of k in Equation (3.35) is close to unity
and may be considered constant at each step without loss of accuracy. On convergence,
p may be calculated and a new k evaluated suitable for the next step.
In Mode 3 operation, the value of k becomes more significant and for this reason
the number of variables is increased to six to include the commutation angle p. The
equations [ F ( X ) ] = 0 for the Newton-Raphson method in this case are:

VLp - VtemL8 - (8.92)

Jz
-uVterm
A
cOs(8 - @ ) f ( p )- -aVterm cosa!' + -3Inxd c = 0, (8.93)
7I 7I
Id - ka * aVtermcos CY'- kb = 0, (8.94)

cos(a!
AXc
+ p + 30) - cos(Y' + -Idavterm = 0, (8.95)

a! - CY,ln = 0. (8.96)
Although k can be calculated explicitly, a linearized form of Equation (8.60) obtained
for 01 = 30" can be used to simplify the expression. In the range 60" < 1 < 120", the
value of k can be obtained from:
= 1.01 - 0.0573 1,
f(1) (8.97)
where p is measured in radians.
In Mode 4, the a s . and d.c. systems are both short circuited at the rectifier and
operate independently. In this case, the system equivalent of Figure 8.34 reduces to
that shown in Figure 8.35. The network equivalent can be solved directly and the d.c.
current is obtained from the algebraic form of the differential Equation (8.57).

Dynamic
element

Figure 8.35 Rectifier load equivalent in Mode 4 operation


8.11 STATIC VAR COMPENSATION SYSTEMS 339

d.c. links The problems associated with dynamic rectifier loads do not occur when
the d.c. link is represented by a quasi steady-state model. Each converter behaves
in a manner similar to that of a converter for a static rectifier load. A nominal bus
shunt admittance (Lo) is calculated from the initial load flow for both the rectifier and
inverter ends and injected currents are used at each step in the solution to account for
the change from steady state calculated from Equation (8.76). Note that the steady-
state shunt admittance at the inverter (YOoi)will have a negative conductance value as
power is being supplied to the network. This is not so for a synchronous or induction
generator as the shunt admittance serves a different purpose in these cases.

8.10.4 Inclusion of converters in the transient stability


program
A flow diagram of the unified algorithm is given in Figure 8.36. It is important to note
that the hyperplanes of the functions used in the Newtonian iterative solution process
are not linear and good initial estimates are essential at every step in the procedure. A
common problem in converter modelling is that the solution converges to the unrealistic
result of converter reactive power generation. It is, therefore, necessary to check against
this condition at every iteration. With integration step lengths of up to 25 mS, however,
convergence is rapid.

8.11 Static VAR Compensation Systems


Static VAR compensation systems (SVS) of large ratings are now in common use to
control the voltage at points remote from power generation. They are also installed to
assist system stability and, therefore, to consider them as a fixed shunt element can
produce erroneous information in a transient stability study.
A composite static VAR compensator, shown in Figure 8.37, has been suggested by
a CIGRE Working Group [16]. The model is not overly complex as this would make
data difficult to obtain and would be incompatible with the overall philosophy of a
multi-machine transient stability program. The SVS representation can be simplified
to any desired degree, however, by a suitable choice of data.
The basic control circuit consists of two lead-lag and one lag transfer function
connected serially. The differential equations describing the action of the control circuit
with reference to Figure 8.37 are:

Although electronically produced, the dead band may be considered as a physical


linkage problem, as shown in Figure 8.38(a). In this example, the input ( x ) and output
( y ) move vertically. The diagram shows the initial steady-state condition in which x
and y are equal. The input ( x ) may move in either direction by an amount Db/2 before
y moves. Beyond this amount of travel, y follows x, lagging by &/2 as depicted in
Figure 8.38(b). The effect of a dead band can be ignored by setting Db to zero.
340 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

TS programme Rectifier subroutines

Calculate constants for


differential equation

( Chec for switching } I .a-


1
I I No Obtain new 1
a 1
Extract th_eveninvoltag-
Thevenin imDendancd

es from [ V ]

II

,
L

t Calculate new initial


vuT
Calculate K, if mAAn lconditons and mode 1
I operation I

3
I v
I Calculate mismatches
I
I Convergence Yes
I
I
I
I
I I
given mode of operation

(-%
I
i lter +
I 11ter+ll

I
I
I Calculate rectifier variables

I t
I
II - lter > 15
+
Yes
)
I
I Error messages I A.C. current on [ V ]
I
!
#
.
(Convergence?) I I
I
Figure 8.36 Unified algorithm flow diagram

Stepped output permits the modelling of SVS when discrete capacitor (or inductor)
blocks are switched in or out of the circuit. It is usual to assume that all blocks are of
equal size. During the study, the SVS operates on the step nearest to the control setting.
Iterative chattering can occur if the control system output (B3) is on the boundary
between two steps. The simplest remedy is to prevent a step change until B3 has
moved at least 0.55BSepfrom the mean setting of the step.
8.1 1 STATIC VAR COMPENSATION SYSTEMS 341

Initial
controlling

Y
vohage ( p u ~ "9v-t
Initial
susceptanm (pu)

Dead band Control system Susceptance


limits and steps

Figure 8.37 Composite static VAR compensation system (SVS) model

(a) Movement
1 x-r+z4
Movement
input x
\n

Figure 8.38 Dead band analogy and effect. (a) Physical analogy of dead band; (b) The effect
of a dead band on output

The initial MVAR loading of the SVS should be included in the busbar loading
schedule data input. However, it is possible for an SVS to contain both controllable
and uncontrollable sections (e.g. variable reactor in parallel with fixed capacitors or
vice versa). It is the total MVAR loading of the SVS which is, therefore, included in
the busbar loading. Only the controllable part should be specified in the SVS model
input and this is removed from the busbar loading leaving an uncontrollable MVAR
load which is converted into a fixed susceptance associated with the network.
Note that busbar load is assumed positive when flowing out of the network. The
sign is, therefore, opposite to that for the SVS loading.
In order to clarify this, consider an overall SVS operating in the steady state, as
shown in Figure 8.39(a). The busbar loading in this case must be specified as -50
MVAR and it may be varied between -10 MVAR and -80 MVAR provided the
voltage remains constant.
The SVS may be specified in a variety of ways, some more obvious than others,
the response of the system being identical. Three possible specifications are given in
Table 8.4. In the first example, the network static load will be +20 MVAR, while in
the second case the static load will be zero. The third example may be represented by
an overall SVS, as shown in Figure 8.39(b).
342 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

(limits
70MVAR0 and

70 MVAR)
12z 80 MVAR

Figure 8.39 Example of an overall SVS controllable and uncontrollable sections. (a) Example
of overall SVS using controllable capacitors; (b) Alternative to overall SVS in (a) using a
controllable reactor

Table 8.4 Examples of MVAR loading specification for SVS shown in


Figure 8.39(a)
Example Initial loading Maximum limit Minimum limit
1 70 100 30
2 50 80 10
3 -30 0 -70

It is convenient, when specifying the initial steady-state operation, to use MVAR.


However, this is a function of the voltage and, hence, all MVAR settings must be
converted to their equivalent per unit susceptance values prior to the start of the stability
study.

8.11.1 Representation of SVS in the overall system


The initial MVAR loading of the SVS is converted into a shunt susceptance (Bo) and
added to the total susceptance at the SVS terminal busbar. During the system study, the
deviation from a fixed susceptance device is calculated ( B 4 ) and a current equivalent
to this deviation is injected into the network.
A reduction in controlling voltage VSV will cause the desired susceptance B4 to
increase. That is, the capacitance of the SVS will rise and the MVAR output will
increase.
The injected current (Ti,,,) into the network is given by:

(8.101)
8.12 REFERENCES 343

where
Y =0 +jB4.
Although not necessary for the solution process, the MVA output from the SVS into
the system is given by:

(8.102)

8.12 References
1, Kulicke, B. Netomac digital program for simulating electromechanical and electromagnetic
transient phenomena in a x . systems, (198 1). Siemens Aktienngesel-lschaft, E15/1722- 101.
2. Woodford, D A, (1985). Validation of digital simulation of d.c. links, IEEE Transactions on
Power Apparatus and Systems, PAS-104 (9), pp. 2588-2595.
3. Anderson, G W J, Arnold, C P, Watson, N R and Arrillaga, J, (1995). A new hybrid
a.c. -d.c. transient stability program, International Conference on Power Systems Transients
(IPST), pp. 535-540.
4. Anderson, G W J, (1995). Hybrid simulation of a.c.-d.c. power systems, Ph.D. thesis,
University of Canterbury, New Zealand.
5 . Eguiluz, L I and Arrillaga, J, (1995). Comparison of power definitions in the presence of
waveform distortion, International Journal of Electrical Engineering Education, 32 (2),
pp. 141-153.
6. Heffernan, M D, Turner, K S, Arrillaga, J and Arnold, C P, (1981). Computation of
a.c. -d.c. system disturbances -Parts I, I1 and 111, Transactions on Power Apparatus and
Systems, PAS-100 (1 l), pp. 4341 -4363.
7. Szechtman, M, Weiss, T and Thio, C V, (1991), First benchmark model for HVDC control
studies, ELECTRA (133, pp. 55-75.
8. Clarke, E, (1950). Circuit Analysis of a.c. Power Systems- Vol. 11, John Wiley and Sons,
New York.
9. Kimbark, E W, (1968). Power System Stability- Vol. IIIs Synchronous Machines, Dover
Publications, New York.
10. Reeve, J and Adapa, R, (1988). A new approach to dynamic analysis of a.c. networks
incorporating detailed modelling of d.c. systems -Parts I and 11, IEEE Transactions, PD-3
(4), pp. 2005-2019.
1 I . Watson, N R, (1987). Frequency-dependent a.c. system equivalents for harmonic studies
and transient converter simulation, Ph.D. thesis, University of Canterbury, New Zealand.
12. Arnold, C P, Turner, K S and Arrillaga, J, (1980). Modelling rectifier loads for a multi-
machine transient-stability programme, IEEE Transactions on Power Apparatus and Systems,
PAS-99 (l), pp. 78-85.
13. Giesner, D B and Arrillaga, J, (1970). Operating modes of the three-phase bridge converter,
International Journal of Electrical Engineering Education, 8, pp. 373-388.
14. IEEE Working Group on Dynamic Performance and Modeling of d.c. Systems, (1980).
Hierarchical structure.
15. Turner, K S , (1980). Transient stability analysis of integrated a.c. and d.c. power systems,
Ph.D. thesis, University of Canterbury, New Zealand.
16. CIGRE Working Group 31-01, (1977). Modelling of static shunt VAR systems for system
analysis, ELECTRA (Sl), pp. 45-74.

Das könnte Ihnen auch gefallen