Chapter 13 Transient Stability
Chapter 13 Transient Stability
Transient Stability
CCT 1
Et
@Jim
X2
Xtr
XI
Chap. 1
b,
/
5Infinite
5 bus
/
CCT 2
the transient reactance (Xi)is denoted by E'. The rotor angle 6 represents the angle
by which E leads EB. When the system is perturbed, the magnitude of E' remains
constant at its predisturbance value and F changes as the generator rotor speed
deviates from synchronous speed we.
The system model can be reduced to the form shown in Figure 13.2(b). It can
be analyzed by using simple analytical methods and is helpful in acquiring a basic
understanding of the transient stability phenomenon. This model is identical to that
shown in Figure 12.4 of Chapter 12. From Equation 12.71, the generator's electrical
power output is
Set. 13.1
rnax
=-
E'EB
,X T
Since we have neglected the stator resistance, P, represents the air-gap power
as well as the terminal power. The power-angle relationship with both transmission
circuits in service (11s) is shown graphically in Figure 13.3 as curve 1. With a
mechanical power input of P,, the steady-state electrical power output P, is equal to
pm, and the operating condition is represented by point a on the curve. The
corresponding rotor angle is 6.,
I-
Transient Stability
830
Chap, 1
The equation of motion or the swing equation (see Chapter 3, Section 3.9)
"ay
be written as
where
P,
P,,
H
6
t
Set, 13.1
,,tor, the rotor angle cannot change instantly from the initial value of 60 to 6,
to the new equilibrium point b at which P, =P, I . The mechanical power
is now in excess of the electrical power. The resulting accelerating torque causes the
rotor to accelerate from the initial operating point a toward the new equilibrium point
b, tracing the P,-6 curve at a rate determined by the swing equation. The difference
between Pmland P, at any instant represents the accelerating power.
When point b is reached, the accelerating power is zero, but the rotor speed
is higher than the synchronous speed wo (which corresponds to the frequency of the
infinite bus voltage). Hence, the rotor angle continues to increase. For values of 6
higher than 61, Pe is greater than P,, and the rotor decelerates. At some peak value
,6, the rotor speed recovers to the synchronous value wo, but P, is higher than P,,,.
The rotor continues to decelerate with the speed dropping below oo; the operating
point retraces the P,-6 curve from c to b and then to a. The rotor angle oscillates
indefinitely about the new equilibrium angle 6 , with a constant amplitude as shown
by the time plot of 6 in Figure 13.4(b).
In our representation of the power system in the above analysis, we have
neglected all resistances and the classical model is used to represent the generator. In
effect, this neglects all sources of damping. Therefore, the rotor oscillations continue
unabated following the perturbation. In practice, as discussed in Chapter 12, there are
many sources of positive damping including field flux variations and rotor amortisseur
circuits. In a system which is small-signal stable, the oscillations damp out.
Equal-area criterion
For the system model considered above, it is not necessary to formally solve
the swing equation to determine whether the rotor angle increases indefinitely or
oscillates about an equilibrium position. Information regarding the maximum angle
excursion (6,) and the stability limit may be obtained graphically by using the powerangle diagram shown in Figure 13.4. Although this method is not applicable to
multimachine systems with detailed representation of synchronous machines, it helps
in understanding basic factors that influence the transient stability of any system.
From Equation 13.3, we have the following relationship between the rotor
angle and the accelerating power:
Transient Stability
Chap. 13
Integrating gives
The speed deviation d6ldt is initially zero. It will change as a result of the disturbance.
For stable operation, the deviation of angle 6 must be bounded, reaching a maximum
value (as at point c in Figure 13.4) and then chaging direction. This requires the
speed deviation d6ldt to become zero at some time after the disturbance. Therefore,
from Equation 13.6, as a criterion for stability we may write
where 60 is the initial rotor angle and 6, is the maximum rotor angle, as illustrated
in Figure 13.4. Thus, the area under the function P, -P,plotted against 6 must be zero
if the system is to be stable. In Figure 13.4, this is satisfied when area A , is equal to
area A2. Kinetic energy is gained by the rotor during acceleration when 6 changes
from 60 to 6,. The energy gained is
61
El
/(pm-pe)d6 = area A,
60
As we have not considered any losses, the energy gained is equal to the energy lost;
therefore, area Al is equal to area A2. This forms the basis for the equal-area criterion.
It enables us to determine the maximum swing of 6 and hence the stability of the
system without computing the time response through formal solution of the swing
equation.
sec, 13.1
833
834
Transient Stability
Chap. 13
d and follows the P,-6 curve for the postfault system farther down. The minimum
value of 6 is such that it satisfies the equal-area criterion for the postfault system. I,
the absence of any source of damping, the rotor continues to oscillate with constant
amplitude.
S ~ C .13.1
835
(b)
The generator output during the fault. This depends on the fault location and
type.
(c)
(d)
(e)
The generator reactance. A lower reactance increases peak power and reduces
initial rotor angle.
(f)
The generator inertia. The higher the inertia, the slower the rate of change in
angle. This reduces the kinetic energy gained during fault; i.e., area At is
reduced.
(g)
The generator internal voltage magnitude (E'). This depends on the field
excitation.
(h)
836
Transient Stability
Chap. 13
where x is the state vector of n dependent variables and t is the independent variable
(time). Our objective is to solve x as a function of t, with the initial values of x and
t equal to xo and to, respectively.
In this section we provide a general description of numerical integration
methods applicable to the solution of equations of e above form. In describing these
methods, without loss of generality, we will treat quation 13.10A as if it were a firstorder differential equation. This simplifies presentation and makes it easier for a
novice reader to comprehend the special features of each method.
We will first describe the Euler method, which by virtue of its simplicity
serves as a good introduction to numerical integration, and then we discuss more
advanced methods.
with x=xo at t=to. Figure 13.6 illustrates the principle of applying the Euler method.
Tangent
Figure 13.6
Sec. 13.2
837
At x=x,, t=to we can approximate the curve representing the true solution by
its tangent having a slope
Therefore,
The Euler method is equivalent to using the first two terms of the Taylor series
expansion for x around the point (x,, to):
After using the Euler technique for determining x=xl corresponding to t=tl, we can
take another short time step At and determine x2 corresponding to t2=t,+At as follows:
Transient Stability
Chap. 13
The standard Euler method results in inaccuracies because it uses the derivative
at the beginning of the interval as though it applied throughout the interval. ~h~
modified Euler method tries to overcome this problem by using the average of the
derivatives at the two ends.
The modified Euler method consists of the following steps:
(a)
Predictor step. By using the derivative at the beginning of the step, the value
at the end of the step is predicted
(b)
Corrector step. By using the predicted value of xf,the derivative at the end of
the step is computed and the average of this derivative and the derivative at
the beginning of the step is used to.find the corrected value
63'
If desired, a more accurate value of the derivative at the end of the step can be
calculated, again by using x=xf. This derivative can be used to calculate a more
accurate value of the average derivative which is in turn used to apply the corrector
step again. This process can be used repeatedly until successive steps converge with
the desired accuracy.
The modified Euler method is the simplest of predictor-corrector (P-C)
methods. Among the well known higher order P-C methods are the Adarns-Bashforth
method, Milne method, and Hamming method [4]. The applicability of these methods
to power system stability analysis has been investigated in reference 7 and has been
found to suffer from a number of limitations. They are not self-starting, need more
computer storage and require smaller time steps than the Runge-Kutta methods
described below.
13.2.3 Runge-Kutta (R-K) Methods [4,5]
The R-K methods approximate the Taylor series solution; however, unlike the
formal Taylor series solution, the R-K methods do not require explicit evaluation of
derivatives higher than the first. The effects of higher derivatives are included by
several evaluations of the first derivative. Depending on the number of terms
effectively retained in the Taylor series, we have R-K methods of different orders.
Set. 13.2
839
where
This method is equivalent to considering first and second derivative terms in the
Taylor series; error is on the order of ~ t ~ .
A general formula giving the value of x for the (n+l)Ststep is
where
The general formula giving the value of x for the (n+l)st step is
where
kl
= f(xn,tn)At
840
Transient Stability
Chap. 13
With xo as the initial value of x at the begimng of a step and by using j=1,2,3
and 4 to denote four stages, each stage of the Gill method can be described as follows
141:
Solution at the end of a time step is given by x4. Initially qo=0, thereafter in advancing
the solution, qo for the next step is equal to q4 of the previous step.
The following are the advantages of Gill's version of the R-K method:
(a)
(b)
Storage requirements are less than for the original R-K method.
Set, 13.2
The accuracy of results obtained with the above numerical integration methods
may be checked by using Richardson's formula which gives the accumulated
@*ncation error propagated in the coyse of integration. The difference between the
true value of a variable and the value obtained by a fourth-order Runge-Kutta method
using a step length of At is given by [4,7]
where
true)
x(At)
x(2At)
true value of x
= value computed with a step length of At
= value computed with a step length of 2At
=
842
Transient Stability
Chap. 13
= f(x,t)
with x=xo at t = t o
Figure 13.7
Set. 13.2
We see that x, appears on both sides of Equation 13.2 1. This implies that the
variable x is computed as a function of its value at the previous time step as well as
the current value (which is unknown). Therefore, an implicit equation must be solved.
The trapezoidal rule is numerically A-stable [ 6 ] . The stiffness of the system
being analyzed affects accuracy but not numerical stability. With larger time steps,
high frequency modes and fast transients are filtered out, and the solutions for the
modes is accurate. For systems involving simulations in which time steps are
limited by numerical stability considerations rather than accuracy, implicit methods
are generally better suited than explicit methods.
The trapezoidal rule is a second-order method. Implicit integration methods of
higher order have been proposed in the literature on numerical methods. However,
they have not been widely used for power system applications since they are more
difficult to program and less numerically stable than the trapezoidal rule.
In the above description of different numerical integration methods, for
simplicity we have considered a first-order differential equation. When applied to the
analysis of power system stability, the system equations are organized as a set of firstorder differential equations. The rate of change of each state variable depends on other
state variables and is not an explicit function of time t. The following example serves
as a simple illustration of the application of numerical integration as well as the equalarea criterion for transient stability analysis.
+
Example 13.1
In this example, we examine the transient stability of a thermal generating station
consisting of four 555 MVA, 24 kV, 60 Hz units supplying power to an infinite bus
through two transmission circuits as shown in Figure E13.1. This system is the same
as the one considered in Example 12.2 of Chapter 12 in which we examined the
small-signal performance.
LT Trans.
4
7.
Y -
jw-
CCT 2
Infinite
5 bus
?5 EB
j0.5
,F
CCT 1
j0.93
QFigure E13.1
Transient Stability
Chap. 13
The network reactances shown in the figure are in per unit on 2220 MVA, 24 kV base
(referred to the LT side of the step-up transformer). Resistances are assumed to be
negligible.
The initial system-operating condition, with quantities expressed in per unit on 2220
MVA and 24 kV base, is as follows:
P=0.9
Q=0.436 (overexcited)
&=I .0~28.340 J ! ? ~ = o . ~ o o ~ ~ L o
Circuit 2 experiences a solid three-phase fault at point F, and the fault is cleared by
isolating the faulted circuit.
(a)
Determine the critical fault-clearing time and the critical clearing angle by
computing the time response of the rotor angle, using numerical integration.
(b)
Check the above value of critical clearing angle, using the equal-area
criterion.
Solution
With the generator represented by the classical model, the system-equivalent circuit
is as shown in Figure E13.2.
13.2
Numerical 1ntegration)Methods
845
Figure E13.3 shows the reduced equivalent circuits representing the three system
conditions: (i) prefault, (ii) during fault, and (iii) postfault. Also shown in the figure
are the corresponding expressions for the electrical power output as a function of 6.
(a) Prefault
(c) Postfault
Figure E13.3 Reduced equivalent circuits and equations for power output
Transient Stability
346
where
lo
1.351
p~
1.1024
Chap. 13
The initial values of 6 and Amr are 41.77" and 0 pu, respectively.
Any of the numerical integration methods described in Section 13.2 may be used to
solve Equations E13.1 and El 3.2. For illustration, let us consider the second-order RK method. The general formulas giving the values of Am, 6, and t for the (n+lyt step
of integration are as follows:
where
Figure E13.4 shows plots of 6 as a function of time, for the three values of faultclearing time (t,): 0.07 s, 0.086 s, and 0.087 s. The corresponding values of clearing
angle (6,) are 48.58", 52.04", and 52.30, respectively. These results were computed
using a time step (At) of 0.05 s throughout the solution. The time step was, however,
adjusted near the fault-clearing time so as to give the exact switching instant.
From the results, we see that the system is stable with tc=0.086 s (6,=52.04'), and is
unstable with tc=0.087 s (6,=52.30); the critical clearing time is, therefore,0.0865-tO.0005 s, and the critical clearing angle is 52.17"~0.13".
13.2
847
The power-angle diagrams for the three network conditions are shown in Figure
E13.5. For the critically stable case, the maximum swing in 6 is given by
1.1024 sin6, =0.9
hence, 6, = 125.27"
fault
848
Transient Stability
Chap. 1
A,
or
set. 13.3
C--------------------------------------------------------------------
i
I
Transmission
network
[-.--.-------1. . . . . . . . . . . . . . . . . . . . . . . . .
**
Individual machine
reference fi-ame: d-q
) Other generators
I
I
equations
/ } Motors
i
:
1 1
'
including
static loads
governor
849
1I
Other dynamic
} devices, e.g.,
HVDC, svc
Common reference
fiame: R-I
* Algebraic equations
** Differential equations
Figure 13.8 Structure of the complete power system model
for transient stability analysis
The model used for each component should be appropriate for transient stability
analysis, and the system equations must be organized in a form suitable for applying
numerical methods.
As we will see in what follows, the complete system model consists of a large
set of ordinary differential equations and large sparse algebraic equations. The
transient stability analysis is thus a differential algebraic initial-value problem.
13.3.2 Synchronous Machine Representation [8]
Transient Stability
Chap,
where
w,
Am,
p
With rotor currents expressed in terms of rotor and mutual flux linkages (see
Equation 12.159 of Chapter 12), the rotor circuit dynamic equations are
set, 13.3
85 1
where
Lags
'19
'29
Here La& and Lagsare saturated values of the d- and q-axis mutual inductances given
by
and K, and Ksq are computed as a function of the air-gap flux linkage
described in Chapter 3 (Section 3.8.2).
vatas
852
Transient Stability
Chap.
with
Since we have neglected the effect of speed variations on the stator voltage,
6 =a/ao= 1.0 in the above equations. Consequently, 6L i =Xiand 615;' =X:. The above
equations are in the individual machine d-q referenc@frame which rotates with the
machine's rotor. For the solution of the interconnecting transmission network
equations, a synchronously rotating common R-I reference is used. The relationships
shown in Figure 13.10 are used to transform variables from one reference frame to
the other. The R-axis of the common reference frame also serves as the reference for
measuring the rotor angle 6 of each machine.
Set. 13.3
RII
X,
xicos26+x:sin26
X,
xisin26+x"cos26
4
In this case, E i +jE; represents the voltage behind the subtransient impedance R, + j Y .
For network solution, the generator may be represented by either of the simple
equivalent circuits shown in Figure 13.1 1 .
854
Transient Stability
Chap.
The air-gap torque required for the solution of the swing equation (13.22) is
Since we have assumed 6 =a/ao= 1.O pu in the ststor voltage equations, in per unit the
air-gap torque is equal to the air-gap power (see Chapter 5, Section 5.1.2). Hence,
The per unit exciter output current Ifd (see Chapter 8, Section 8.6) is
Here we have considered a generator model with one d-axis and two q-axis
amortisseur circuits. For models with a different number of rotor circuits, changes to
the above formulation of machine equations are straightforward. However, we have
assumed equal mutual inductances between the armature and rotor circuits in each
axis. Reference 8 provides a description of the implementation of a model with
unequal mutual inductances for stability analysis, using the above general approach.
set 13.3
855
1
* - .v1
l+sT,
Gain
Washout
Exciter
Phase
compensation
vs
VS max
E~rnax
856
Transient Stability
Chap.
For a bus-fed (potential source) thyristor exciter, the voltages vary with the
generator terminal voltage (E,) and exciter output current (I/,):
with
with
set. 13.3
857
Both limits considered here are windup limits. Modelling of non-windup limits
is discussed in Chapter 8.
The generator field voltage efd in the reciprocal per unit system is related to
exciter output voltage Efd (see Chapter 8, Section 8.6) as follows:
~n alternative method of treating blocks 4 and 5, in which the derivatives of the input
variables are not required, is described in Chapter 12, Section 12.3.4.
Initial values of excitation system variables:
For any given steady-state generator output, the field voltage efd is determined
by the generator equations (see Section 13.3.2). The excitation system quantities are
determined as follows:
Thus Vmf takes a value appropriate to the generator loading condition prior to the
disturbance.
Transient Stability
Chap.
The transients associated with the transmission network decay very rapidly.
fact, network transients will have died out by the time the solution of the swing
equation is advanced by one time step. Therefore, it is usually adequate to regard the
network, during the electromechanical transient conditions, as though it were passing
directly from one steady state to another. The fundamental frequency (50 or 60 H ~ )
variations may be regarded as microprocesses, and only the variations in the envelopes
(amplitude modulation) of current and voltage waveforms are considered for stability
analysis. For analysis of balanced conditions, a single-phase representation of the three
phases is used. Unbalanced faults are simulated by using symmetrical components as
described in Section 13.4.
Without the balanced steady-state representation of the transmission network,
stability analysis of large practical power systems would be impractical. For special
electromechanical problems requiring inclusion of transmission %&work and generator
stator transients, and three-phase representation, a program such as the electromagnetic
transients program (EMTP) may be used [9]. For conventional transient stability
analysis, the network representation is similar to that for power-flow analysis. As
described in Chapter 6 (Section 6.4.1.), the most convenient form of network
representation is in terms of the node admittance matrix. The structure and
formulation of this matrix are presented in Section 6.4.1.
The characteristics and modelling of loads are covered in Chapter 7. Dynamic
loads are represented as induction or synchronous motors, and their treatment is
similar to that of synchronous machines.
Static loads are represented as part of the network equations. Loads with
constant impedance characteristics are the simplest to handle and are included in the
node admittance matrix. Nonlinear loads are modelled as exponential or polynomial
functions of bus voltage magnitude and frequency. The net effect is that a nonlinear
static load model is treated as a current injection at the appropriate node in the
network equation. The value of the node current from ground into the network is
vi
where
is the conjugate of the load bus voltage, and PL and QL are portions of the
active and reactive components of the load which vary as nonlinear functions of VL
and frequency deviation. For an inductive load QL is positive.
The overall networkfload representation comprises a large sparse nodal
admittance matrix equation with a structure similar to that of the power-flow problem
discussed in Chapter 6, Section 6.4. The network equation is identical to Equation
6.87, which in matrix notation may be written as
<
set. 13.3
859
Simulation of faults:
A fault at or near a bus is simulated by appropriately changing the selfadmittance of the bus.
Methods of simulating different types of fault will be discussed in Section
13.4.5. Depending on the type of fault, the equivalent negative- and zero-sequence
impedances seen at the fault point are computed, combined appropriately, and inserted
between the fault point and ground. This simply alters the self-admittance of the node
representing the fault bus.
For a three-phase fault, the fault impedance is zero andcihe faulted bus has the
same potential as the ground. This involves placing an infinite shunt admittance. In
practice, a sufficiently high shunt admittance (i.e., a very small shunt impedance) is
used so that the bus voltage is in effect zero. For example, a conductance G equal to
lo6 pu on 100 MVA base would effectively reduce the bus voltage to zero. The fault
is removed by restoring the shunt admittance to the appropriate value depending on
the postfault system configuration.
13.3.5 Overall System Equations
Transient Stability
Chap.
where
state vector of individual device
Id = R and I components of current injection from the device into the network
Vd = R and I components of bus voltage
xd
The overall system equations, including the differential equations (13.5 1) for
all the devices and the combined algebraic equations for the devices (13.52) and the
network (13S O ) are expressed in the following general form comprising a set of firstorder differential equations
Time t does not appear explicitly in the above equations. A wide range of approaches
has been reported in the literature for solving these equations, depending on the
numerical methods and modelling details used. Reference 10 provides a review of
these approaches. The many possible schemes for the solution of Equations 13.53 and
13.54 are characterized by the following factors:
(a)
The manner of interface between the differential equations 13.53 and the
algebraic equations 13.54. Either a partitioned approach or a simultaneous
approach may be used.
(b)
(c)
The method used for solving the algebraic equations. As in the case of powerflow analysis described in Chapter 6, one of the following methods may be
used: (i) the Gauss-Seidal method based on admittance matrix formulation, (ii)
set. 13.3
861
In this approach, the algebraic and differential equations are solved separately.
Initially, at t=O-, the values of the state variables x and the network variables V and
I are known, and they form a consistent set; that is, the system is in steady state and
the time derivatives f(x,V) are equal to zero.
Following a disturbance, usually a network fault, the state variables x cannot
change instantly. The algebraic equations 13.54 are solved first to give V and I, and
the corresponding power flows and other non-state variables of interest at t=O'. Then
the time derivatives f(x,V) are computed by using the known values of x and V in
Equations 13.53. These then can be used to initiate the solution of the state variables
x by using any of the explicit integration methods described in Section 13.2.
We will illustrate this by considering Gill's version of the fourth order R-K
method. There are four stages (j=1 to 4) per time step At. Using the values of f(x,V)
computed at the beginning of a time step, k,, x,, and q, are computed according to
Equations 13.17 with j= 1. Then the algebraic equations 13.54 are solved with x =a,to
compute V1 and I,. These are in turn used in Equations 13.17 to compute k2, x2, and
q,. This process, involving alternating solutions of algebraic and differential equations,
is applied successively with x4 representing the solution at the end of each time step.
During a switching operation, the network variables change instantly, but not the state
variables.
Since the solution of differential equations requires values of network and state
variables only from the previous steplstage, the differential equations may be
partitioned in any way desired. For example, differential equations associated with
each device may be solved independently. This offers considerable programming
flexibility.
For solution of networks associated with large interconnected systems, the most
efficient method is to use sparsity-oriented triangular factorization.
The partitioned approach with explicit integration is the traditional approach
used widely in production-grade stability programs. Its advantages are programming
flexibility and simplicity, reliability, and robustness. Its principal disadvantage is
susceptibility to numerical instability. For a stiff system, a small time step is required
throughout the solution period, as dictated by the smallest time constant (or
eigenvalue).
862
Tran.sient Stability
Chap. 1
In this approach, the state variables and the network variables are solved
simultaneously. We will illustrate this using the trapezoidal rule.
With X=H, and V=Vn at t=tn, the solution of x at t=t,+, =tn+Atis given by
applying the trapezoidal rule (Equation 13.21) to solve Equation 13.53:
The vectors
and
At solution,
Equations 13.59 and 13.60 are both nonlinear algebraic equations. Thus the
differential equations have been made algebraic by using an implicit formula. These
equations are very sparse; for computational efficiency it is necessary to take '
advantage of their special structure. Applying the Newton method (see Chapter 6) to
solve Equations 13.59 and 13.60, we may write for the (k+l)stiteration
Set. 13.3
v = v ~ ,It+has
~ .the
The matrices AD, BD, CD and YD are associated with the models for the dynamic
devices and nonlinear static loads. For a system with m such devices, they have the
following structures:
The solutions of Equations 13.59 and 13.60 are given in terms of the above matrices
by
In the above equations, k is the iteration counter and good starting values
v:+~)
are established by extrapolation. Also F:+~ and ~ k , , , are the residue vectors of the
states and current injections, respectively.
From Equation 13.64,
:
can be expressed as a function of AV~,+,
Transient Stability
Chap,
Now AV:, and AX:,, can be calculated by solving Equations 13.66 and 13.67. Then
and v!:: are obtained from Equation 13.61.
$:
Treatment of discontinuities:
Equations 13.64 and 13.65 are valid only when the functions given by
Equations 13.57 and 13.58 are continuous and differentiable. At points of
discontinuity, such as network switching or limits on state variables, the exact
formulation by the Newton method or any method requiring derivatives would be
complicated [13]. This problem is dealt with in reference 14 as @llows:
For large network discontinuities, such as a network fault or switching
operations, the integration method is temporarily changed to the fourth order
Runge-Kutta method for one step at the point of discontinuity. This time step
has a zero step size and is used only for the calculation of the postfault
network conditions (the state vector is not updated). After this, the normal
trapezoidal integration is resumed.
For local non-differentiable functions, such as limits associated with
controllers, the device Jacobians are computed by neglecting their effect. This
is acceptable since they have only local impact and the overall convergence
will not be significantly affected.
Example 13.2
In this example, we analyze the transient stability of the system of Figure E13.1
(considered in Example 13.1) including the effects of rotor circuit dynamics and
excitation control. The system diagram is reproduced here as Figure E13.6 for,
reference.
Infinite
bus
EB
Network reactanc
Figure E13.6
13.3
865
Generator parameters:
The four generators of the plant are represented by an equivalent generator whose
parameters in per unit on 2220 MVA base are as follows:
0.031
B,,,
6.93
WTI
0.8
Vn=
KA = 200
KmB = 9.5
Vs m m = 0.2
TR = 0.015 s
Tw = 1.41 s
Vsmin = -0.2
EFmm = 7.0
TI = 0 . 1 5 4 ~
EFmin= -6.4
T2 = 0.033 s
0.9
= l.OL28.34O
El
0.436 (overexcited)
EB = 0.90081LO
=
Disturbance:
A three-phase fault on circuit 2 at point F cleared by isolating the faulted circuit
simultaneously at both ends.
Examine the stability of the system with the following alternative forms of excitation
control:
(0
(ii)
(iii)
Tran.sient Stability
Chap. 1
0.07 s
(b)
0.10s
Solution
The generators of this example have the same characteristics as the generator
considered in examples of Chapters 3 and 4.
The per unit fundamental parameters as determined in Example 4.1 are as follows:
The initial per unit values of generator variables based on the results of Examples 3.2
and 3.3 are as follows:
The initial per unit values of exciter output current and voltages are
With EB assumed to be in phase with the R-axis, the initial value of the rotor angle
6,+28.34"
39.1+28.34"
67.44"
%
3.3
180
II
V1
a>
120 -
.H
CQ
hl)
2
8
60 :
Prefault
L'
' I I'
\
I
I
Oo
2
3
Time t in seconds
Figure E13.7 (a) Rotor angle response with fault cleared in 0.07
Figure E13.7 (b) Active power response with fault cleared in 0.07
Transient Stability
Chap.
--..---------I
0.8 I
0.6 0.4 -
Constant Efd
\
1
t
I
I
,\
"/
0.2 0
2
3
Time t in seconds
Figure E13.7 (c) Terminal voltage response with fault cleared in 0.07 s
Time t in seconds
Figure E13.7 (d) Exciter output voltage response with fault cleared in 0.07 s
Set. 13.3
869
With the PSS, the rotor oscillations are very well damped without compromising the
first-swing stability.
The results with regard to oscillatory stability (i.e., damping of rotor oscillations) are
consistent with those of Examples 12.4 and 12.5 of Chapter 12.
(b) Transient response with the fault-clearing time t, equal to 0.1 s:
The transient responses of rotor angle 6 with the three alternative forms of excitation
control are shown in Figure E13.8. With constant Ef, the generator is first-swing
unstable. With a fast-acting exciter and AVR, the generator maintains first-swing
stability but loses synchronism during the second swing. The addition of a PSS
contributes to the damping of second and subsequent swings.
It is evident from these results that the use of a fast exciter having a high-ceiling
voltage and equipped with a PSS contributes to the enhancement of the overall system
,
stability.
n s t i t E~
\
\
\
I
I
;i /
'I\
i,/
'i
:i
1;
/;
I!
i
I
1
Prefault
I:;
1:
1:
/'
\',
ilk
'.
,--.
,'',
: ,',
:
;
%'
st
:
:
'\
:
:
!,
,*--\
\
\
:;
',
, ,>'
i
,,
/'
%-.'
?.', :i
, ,
:I
I
'..'
:I
\i !I
$,, /;
'!
t,=O.l s
L'
i'
'4
Time t in seconds
Figure E13.8 Rotor angle response with fault cleared in 0.1 s
Example 13.3
For the system considered in Example 13.2, examine the accuracy and numerical
stability of Gill's version of the fourth order R-K method and the trapezoidal rule.
Consider the case with AVR and PSS, and the fault cleared in 0.07 s.
Transient Stability
Chap. 1
Solution
The system considered here includes the effects of rotor circuit dynamics and a fast
exciter with a PSS. This represents a very stiff system as is evident from the
eigenvalues computed in Example 12.5 of Chapter 12 (see Table E12.2).
R-K method:
Figure E13.9 shows plots of rotor angle transient responses computed by using four
different values of time step (At): 0.0 1 s, 0.03 s, 0.04 s, and 0.043 s.
With time steps of 0.01 s and 0.03 s, the results are practically identical. When the
time step is increased to 0.04 s, errors become noticeable. With a step size of 0.043 s,
the solution is numerically unstable.
In Table E13.1, the accuracy of results obtained with the R-K m@hod is checked by
applying Richardson's formula (see Equation 13.18) to the results obtained with step
sizes of 0.01 s and 0.02 s. The magnitude of the error is seen never to exceed 0.050,
The table indicates the manner in which the error propagates. It is evident that the
accumulated error oscillates, becoming alternately positive and negative.
01
0
2
3
Time t in seconds
I
5
Figure E13.9 Effect of integration step size on the results of R-K method
13.3
Accumulated error
1
15
Trapezoidal rule:
The results obtained with step sizes of 0.01 s, 0.059 s, 0.09 s, and 0.093 s are shown
in Figure E13.10.
With step sizes of 0.01 s and 0.059 s, the results are practically identical. When the
step size is increased to 0.09 s, significant errors result; however, the general
characteristics of the overall response are still retained. With a step size of 0.093 s,
the errors become so large that the solution blows up after about 1.7 s.
Transient Stability
Chap. 1
Time t in seconds
Figure E13.10 Effect of integration step size on the results of trapezoidal rule
I
set. 13.4
873
(b)
(C)
Figure 13.13 shows the three sets of balanced phasors that form the symmetrical
components of three unbalanced phasors.
(a) Positive-sequence
components
(b) Negative-sequence
components
(denoted by
subscript 1)
(denoted by
subscript 2)
(c) Zero-sequence
components
(denoted by
subscript 0)
Definition of operator a
The operator a denotes 120" phase shift. It is equal to the unit vector ej120 .
Recalling that eje=cosO+isin@,the unit vector representing a is given by
-0.5 +j0.866
The operators a2 and a3 represent phase shifts of 240" and 360, respectively. Thus,
a 2 = e~2400=
and
-0.5-j0.866
874
Transient Stability
Chap.
In matrix form,
Figure 13.14 depicts graphically the synthesis of a set of three unbalanced phasors in
accord with the above equations.
Set, 13.4
similar transformations apply to currents. The sum of the three line currents is equal
to the neutral current. Thus,
Hence,
where Zabc is the impedance matrix giving self- and mutual impedances in and
between phases. ,
By using the transformation of Equation 13.70, Equation 13.74 can be
expressed in terms of the sequence components of voltages and currents:
For typical power systems Zabcis not diagonal but does possess certain symmetries.
These symmetries are such that the sequence impedance matrix ZOl2is diagonal,
either exactly or approximately. Sequence impedances of different types of system
elements will be discussed in Section 13.4.2.
Transient Stability
Chap.
In the above, the currents are line values and the voltages are line-to-neutral values.
It is interesting to note that there are no "cross" terms or components caused by
interaction between voltages of one sequence and currents of another sequence. The
factor 3 is reasonable when we think in terms of 9 components. This also means that
the transformation is not power invariant.
set, 13.4
877
Sequence impedances
The impedance of a circuit when positive-sequence currents alone are flowing
is called the impedance to positive-sequence currents or positive-sequence impedance.
similarly, impedance to negative-sequence currents is called negative-sequence
impedance, and the impedance to zero-sequence currents the zero-sequence
impedance.
A knowledge of the sequence impedances of the power system elements is
essential for the analysis of system performance during unbalanced faults.
1
-
Reference bus
Transient Stability
Chap. 1
Negative-sequence reactance ( X j :
When fundamental frequency negative-sequence (reverse phase sequence)
currents are applied to the armature windings of a machine, with the rotor running at
synchronous speed and the field winding shorted through the exciter, the ratio of the
resulting negative-sequence fundamental frequency voltages to the currents gives the
negative-sequence impedance. The reactive part of this impedance is the negativesequence reactance X2.
In what follows, we will show that the negative-sequence reactance X, is equal
to the mean between the d- and q-axis subtransient reactances, Xz and X:. Saliency
effects result in somewhat different values of X, depending on whether sinusoidal
currents are circulated or sinusoidal voltages are applied.
With time t in seconds, the negative-sequence currents may be expressed as
set. 13.4
879
The resulting field due to the negative-sequence stator currents rotates at synchronous
speed opposite to the direction of the rotor. Therefore, as viewed from the rotor, the
$ator currents appear to be double frequency currents id and i,. Currents of twice the
frequency are induced in the rotor circuits, keeping the flux linkages of the rotor
almost constant (since resistance is considerably smaller than inductance). The
flux due to the stator currents is forced into paths of low permeance which do not link
with any rotor circuit. These flux paths are the same as those encountered in
evaluating subtransient inductances. Therefore, the d- and q-axis components of stator
at 120 Hz encounter subtransient inductances L i and L: (see Chapter 4,
section 4.2). The corresponding per unit stator flux linkages are
Neglecting the stator resistance, with time t in seconds and a, equal to 1.O pu, the per
unit stator voltages are given by the following equations:
.
For fundamental frequency negative-sequence currents, w =ao.Hence,
e,
Similarly,
and
Transient Stability
Chap.
,-
Fundamental frequency
Third harmonic
Similarly,
We see that the voltages have a third harmonic negative-sequence component which
is a function of subtransient saliency (X:-X;).This component is small; unless the
system resonates near this harmonic, it may be neglected. As far as the fundamental
frequency negative-sequence currents are concerned, the reactance is equal to the
average of the subtransient reactances:
Since the rnmf wave due to negative-sequence stator currents moves at twice the
synchronous speed with respect to the rotor, it alternately meets permeances of the
two rotor axes, corresponding to Xi and X;.Hence the value of X2 lies between Xj
and X;. With sinusoidal negative-sequence currents applied, the negative-sequence
reactance is the arithmetic mean of Xi and X;.This would apply when the external
circuit has a large reactance in series.
If sinusoidal negative-sequence voltages are applied to the stator, the resulting
currents in the stator windings can be shown to have a fundamental frequency
component proportional to
set. 13.4
88 1
<").
--Proportional to 2(x,'.
oegative-sequence reactance is
When X$=X;,
there is no third harmonic component. In this case, the voltages
and currents would have only fundamental frequency components, and X,=X$ =X;.
X,
Xr
R,
magnetizing reactance
= rotor leakage reactance
= stator resistance
Xz
Rr
s
882
Transient Stability
Chap,
The real part of the equivalent impedance seen from the stator terminals is the
negative-sequence resistance R2. Since X, is large, it may be neglected. Hence,
Reference bus
Figure 13.17 Negative-sequence equivalent circuit
of a synchronous machine
Zero-sequence impedance
The machine must be star-connected; otherwise, the term zero-sequence
impedance has no significance as no zero-sequence current can flow.
Zero-sequence reactance (Xo):
When zero-sequence currents are applied to the armature, the instantaneous
values of currents in the three phases are equal. With the armature windings infinitely
distributed so that each phase produces a sinusoidal spatial distribution of the rnmf,
the net field produced by the three phases with equal instantaneous currents is zero.
No flux is produced across the air-gap. The only reactance associated with any phase
winding is due to slot leakage flux and end-winding leakage flux. However, in an
actual machine, the winding distribution is not perfectly sinusoidal, and the flux due
to the three phases will have a small value. Therefore, the actual X,is slightly higher
than for the ideal case and depends on pitch and breadth factors of the winding.
Zero-sequence resistance (Rd .
The zero-sequence resistance of a synchronous machine is somewhat larger
than the positive-sequence resistance for reasons discussed below.
set, 13.4
883
The spatial distribution of stator windings is usually not purely sinusoidal, but
of higher harmonics. The spatial third harmonic contribution to the mmf due
to stator currents i,, ib ic is
MMF
K[iacos30+i,cos3(0-1200)+i,cos3(0+1200)]
m e n the stator currents form a balanced set, such as the positive- or negativesequence, the net mmf due to spatial winding distribution harmonics is zero. However,
for zero-sequence currents, with ia=ib=ic=Imcosotand the rotor travelling at
synchronous speed (@=at),the mrnf due to third harmonic spatial distribution is
MMF
3
2
= -I,
Thus the mmf due to third harmonic spatial winding distribution, when zero-sequence
currents flow in the stator windings, consists of second and fourth harmonic
components. These induce currents i~ the rotor and cause rotor heating. Therefore, the
zero-sequence resistance (R,) is slightly higher than the positive-sequence resistance
(RI). The difference is usually not significant.
Figure 13.18(a) shows the flow of zero-sequence currents in the three phases
and the neutral of the synchronous machine. The corresponding (single-phase)
equivalent circuit is shown in Figure 13,18(b).
40
oc
Reference bus0
884
Transient Stability
Chap. 1
The neutral impedance 2, carries a current equal to the sum of the zerosequence currents in the three phases. This is reflected as 32, in the equivalent circuit
of Figure 13.18(b); which represents only one phase. The reference bus for the zerosequence equivalent circuit is the ground at the generator. If the neutral is isolated
from the ground, 2, =aand I,, =O.
13.4.3 Sequence lmpedances of Transmission Lines
See. 13.4
zero-sequence impedances
Transient Stability
Zero-seq uence
equivalent circuit
Winding connection
Case
Chap. 1
1.
0 Ref.
0----------
41
Ref.
lo
o
Figure 13.19 Zero-sequence equivalent of three-phase twowinding and three-winding transformer banks
Ref.
0 Ref.
Sec. 13.4
va = 0
I,
Ic
In terms of the symmetrical components' (see Equations 13.69 and 13.71), we have
va2j I
888
Transient Stability
Chap, 1
The sequence network connection reflecting the above relationships is shown in Figure
13.22(b). Since the fault does not involve ground, the zero-sequence network is absent,
i'Ial
'''a2
xF
xF
set. 1 3.4
889
vI Lq.
F
pos.
net.
Transient Stability
Chap. 13
There are no zero-sequence or negative-sequence emfs, and the voltages V, and Vo2
at the fault point F are each equal to zero. Therefore, zero- and negative-sequence
currents do not flow anywhere in the system. As expected, only the positive-sequence
network is involved. The fault is simulated by shorting the fault point F and the
reference bus R of the positive-sequence network, as shown in Figure 13.24(b).
set. 13.4
891
of fault and inserted as the effective fault impedance Zef in the positive-sequence
Fault type
ValI I
z2z0
2 2 +2 0
Three-phase
This representation gives the correct voltages and currents in the positivesequence network.
There are no negative- or zero-sequence voltages generated in the system. The
negative- and zero-sequence currents that flow during unbalanced faults are driven by
the positive-sequence voltage sources. The positive-sequence power outputs computed
represent the actual power supplied electrically by the machines. This includes all
negative- and zero-sequence resistance losses supplied electrically.
Negative-sequen ce braking torque
892
Transient Stability
Chap. 1
Example 13.4
Figure E13.11 shows the system representation applicable to a 1,000 M V A , 20 kV,
60 Hz generating unit.
HI
~ ' 6
XI=Xo=O.
15
Lbl
X, =0.6,Xo=1.8
F
5I ~ I ~ I I I I L ~ .
2 bus
X,=Xo=O.l
2~ ~
/
CCT 2
5- X , =0.6,Xo=1.8
Figure E13.11
~ 1 . 0
The transmission data shown on the figure are in per unit on 1,000 MVA, 20 kV base.
Network resistances are assumed to be negligible.
The generator data in per unit on the rating of the unit are as follows:
The neutrals of the generator stator and the step-up transformer HT windings are
solidly grounded. The neutral of the star-connected winding of the transformer near
the infinite bus is ungrounded.
-(a)
13.4
893
(i) Neglecting all resistances, find the value of the effective fault impedance
Z,/which, when inserted in the positive-sequence network, represents the
unbalanced fault.
(ii) If the initial generator output conditions are
Solution
For conditions immediately after the fault occurs, the effective positive-sequence
machine reactance is X" = X j=Xg.
The three-sequence networks of the system are shown in Figure E13.12. Since
X j =X," =X,, the negative- and positive-sequence reactances are equal for each of the
elements in the system.
(a) Double line-to-ground fault
(i) Each sequence network represents a balanced system. The unbalanced effect
introduced by the fault at location F is represented by interconnecting the sequence
networks as shown in Figure E 13.13.
XI1=0.25
A- mrm
Gen.
I$&,,
0.6
HI
0.15
r
QTJT-n
Trans.
H2
CCT 1
F
V
0.6
B-
.".. CCT 2
0.1
Trans.
0 2 B
I
--
Transient Stability
Infinite
bus
I
-
$+
5fl
r,',
Fx
Negative
sequence
Fx
Positive
sequence
Zero potential
bus
Zero potential
$
-lr
bus
rr.
Figure E13.13
--.
!i
Fx
Zero
sequence
Zero potential
9'bus
1 3.4
895
Figure E13.14
The effective fault impedance representing the L-L-G fault is
Chap. l 3
Transient Stability
Generator
~
j0.4
j0.4
,-.
Tv
V
=. O m
l OO
j0.0857
j0.2
Infinite bus
+;;;
=>
0 1 . 0 ~ 0 ~
--
The total positive-sequence fault current splits into negative- and zero-sequence
components as shown in Figure E13.15(b).
?-,en.
1.~~7:
Po;itive
sequence
Negative
sequence
sequence
3.5
-
Figure E13.16
13.4
2,
Z2+Zo = j(0.2+0.15)
= j0.35
pu
(ii) Figure E l 3.17 shows current flows in different parts of the sequence networks.
Positive
sequence
~~$-=
1.82
x
F
Negative
sequence
0.9 111
0.91
F
I
1.82
Zero
sequence
Figure E13.17
Transient Stability
Chap. 13
Line-to-line fault
(i) The effective fault impedance is
2,
2,
j0.2
Figure E13.18
TM
(4- R,) I:
Set. 13.4
"a1
'a 1
X.
*Y
Pos. seq. net.
R
52
"a2
'a2
X*
.Y
Neg. seq. net.
R
"a0
The connections of sequence networks so as to satisfy the above equations are shown
in Figure 13.26(b).
(b) Two open conductors
We now consider the condition with phases b and c open between points X
and Y as shown in Figure 13.27(a). The applicable equations are
Transient Stability
Chap,
'a2
X.
.Y
Neg. seq. net.
R
xi bY
Zero seq. net.
R
Val
"a0
__C
Ial.
'a2
X.
.Y
Pos. seq. net.
R
The sequence network connections that satisfy the above equations are shown in
Figure 13.27(b).
From Figure 13.26 we see that, for the one-phase open-circuited condition, the
three-sequence networks are connected in parallel at points X and Y; therefore, even
if the zero-sequence circuit as seen from X and Y has infinite impedance, some power
is transferred through the line between X and Y.
For the case with two phases open-circuited as in Figure 13.27, the threesequence networks are connected in series; unless the zero-sequence circuit provides
a path having finite impedance, no power can be transferred through the line.
Example 13.5
The system shown in Figure El 3.1 1 (Example 13.4) is subjected to a single line-toground fault at point F on circuit 2. The fault is cleared by single-phase switching
which disconnects the faulted conductor at both ends simultaneously. Show how the
network may be represented in stability studies during the period when one conductor
of circuit 2 is switched out.
13.4
Solution
X
I
i-i
I,
0.15
H*
"a
Y
I
Figure E13.19
0.1
CCT 2
0.6
w
77
"t
Positive-sequence network
'a2
0.25
Gen.
CCT 1
0.6
0
0.15
Trans.
0.1
Trans.
CCT 2
0.6
CCT I
Negative-sequence network
1.8
0.15
r'Er'-%z
CCT 2
1.8
CCT 1
Zero-sequence network
Figure E13.20
-3-
0.1
Trans.
/
/
902
Transient Stability
Chap. 13
and the net effective impedance that has to be inserted between points X and Y of the
positive sequence network is
Thus, from the viewpoint of power transfer between the generator and the infinite bus,
the effect of opening one phase of circuit 2 is the same as increasing the reactance of
I
the circuit by 0.7024 pu.
Sec. 13.5
There are a variety of line protection schemes and practices used by utilities
to meet particular system requirements. The following factors influence the choice of
the protection scheme:
Type of circuit: single line, parallel line, multiterminal, magnitude of fault
current infeeds, etc.
Function of line, its effect on service continuity, speed with which fault has to
be cleared
Coordination and matching requirements
Three basic types of relaying schemes are used for line protection: (a)
overcurrent relaying, (b) distance relaying, and (c) pilot relaying [22].
904
Transient Stability
Chap. 1
Impedance relay
Reactance relay
Mho relay
Modified mho and impedance relays, and combinations thereof
Sec. 13.5
(a) Impedance
(b) Reactance
(c) Mho
(f) Lens
Directional
906
Transient Stability
2ndzone
i
I
1St zone
---------------I
Chap. 1
Relay characteristic
at breaker A
Relay characteristic
bat breaker B
,---------d
Distance
a section of the adjacent line CD, they are set to trip after a time delay (typically 0.3
to 0.5 seconds) thereby providing time coordination with the first-zone relays at c
protecting the adjacent line CD.
The third-zone relays at A provide backup protection for the adjacent line CD.
They are set to reach beyond line CD, so that relay operation is assured for faults on
the line. Time coordination with primary protection of line CD is achieved by having
a time delay of about 2 seconds for the third-zone protection.
A protection scheme similar to that at A is provided at B, but set to reach in
the opposite direction. Thus a fault on line AB opens breakers at both ends of the
line.
The number of relays at each end of a line is determined by the requirement
that three-phase, phase-to-phase, phase-to-ground, and double phase-to-ground faults
be covered. Usually, two sets of relays are provided: one set for phase faults and the
other for ground faults.
The reactance type relays are used for short lines, and the mho type relays for
long lines. Lens type relays are used when needed to restrict the tripping area of mho
units for protection of long lines.
Straight-distance relaying provides adequate protection for many situations.
However, it is not satisfactory when simultaneous high-speed tripping of both ends
of the line is critical to maintain system stability. Pilot-relaying schemes, described
below, are better suited for such applications.
(c) Pilot-relaying schemes
Set. 13.5
907
Station A
Z,
station
L
B
~
Transient Stability
Chap.
If the apparent impedances measured by the relaysat both ends are within zone
2 characteristics, then breakers at both ends open at high speed.
If the apparent impedance is within zone 2 characteristics of the relays at only
one station, the breakers at that station are tripped with a time delay of 0.4 seconds,
provided the impedance remains within zone 2 for 0.4 seconds.
To illustrate how this scheme operates, consider the faults F,, F,, and Fj as
shown in Figure 13.32.
Set. 1 3.5
For the fault at F, within the zone 1 characteristics of the relays at both ends,
breakers at the two ends are tripped without intentional time delay. For the fault at F2
within zone 1 characteristic of the relay at B and zone 2 characteristic (but outside
zone 1) of the relay at A, breakers at B trip instantaneously; the zone 2 relays at both
are picked up and they send a.permissive signal to each other. The zone 2 relay
at A, upon receiving permissive signal fiom B (channel time not more than 20 ms),
trips the breakers at A.
For the fault at F3 (outside of the line section), the zone 2 relay at A is picked
up. As the zone 2 relay at B is not picked up, no tripping occurs at B and no
permissive signal is sent to A. The zone 2 relay at A trips the breakers only if the
fault is not cleared by the protective relays of line section BC within 0.4 seconds.
Directional comparison scheme:
The principle of operation of a directional comparison scheme with a blockingtype pilot is illustrated in Figure 13.33. The pilot channel is used to block tripping for
faults external to the protected line.
Each terminal station of the line has three sets of relays:
Underreaching zone 1 phase and ground directional distance relays covering
about 80% of the impedance of the line. These relays operate instantaneously
to trip the local breakers.
Overreaching zone 2 phase and ground directional distance relays set to reach
beyond the remote terminal so as to cover about 120% of the line impedance.
This ensures that all internal faults are detected. These relays trip local
breakers after a delay of about 25 ms (pickup time of the coordinator timer),
if no blocking signal is received from the remote terminal. Zone 2 relays trip
local breakers if the fault is not cleared in about 0.4 s, irrespective of any
blocking signal received from the remote end.
Zone 3 reverse blocking directional distance relays set in a direction opposite
to the protected line. They detect external faults and send a blocking signal to
the remote end to prevent tripping.
Transient Stability
Station A
Station B
Chap,
station c
Set 13.5
91 1
0. However, a local timed trip results, if the fault is not cleared and the zone 2 relay
at station C continues to detect the fault after about 0.4 s.
For a fault at Fj, the zone 2 relays (2;) at station A detect the fault but are
Pgvented from tripping by the blocking signal received from the zone 3 relays (2;')
at
B. Similarly, a fault at F4 is detected by the zone 2 relays (2;) at station B,
but is blocked by signals from the zone 3 relays (2;) at A.
Transient Stability
Bus A
Bus B
Local (bus A)
breakers 1 and 2
Primary relay time
(fault detection)
Auxiliary relay(s) time
Communication time
Breaker trip module
Breaker-clearing time
Chap, 1
Remote (bus B)
breakers 3 and 4
25 ms
3 ms
-
3 ms
33 ms (2 cycles)
Total time
9 ms
17 ms (microwave)
3 ms
50 ms (3 cycles)
64 ms
Notes: (i)
(ii)
Sec. 13.5
Bus A
Bus B
Bus C
Stuck breaker
0
Breaker 4 assumed to be stuck
Breakers 1, 2, 3, 4, and 5 assumed to be 2-cycle air-blast breakers (33 ms)
Breakers 6 and 7 assumed to be 3-cycle oil breakers (50 ms)
Local breaker Remote breakers Local backup Remote backup
6 and 7
breaker 3 breakers 1 and 2
5
Primary relay time
(at bus B)
25 ms
25 ms
25 ms
25 ms
Auxiliary relay(s)
time
3 ms
9 rns
6 ms
12 ms
17 ms
17 ms
3 ms
3 ms
3 ms
3 ms
33 ms (2 cyc.)
5 0 m s (3 cyc.)
33 ms
33 ms
Communication
channel time
Breaker failure
timer setting
Breaker-tripping
module time
Breaker time
--
Total time
64 ms
104 ms
--
157 ms
180 ms
(ii)
Breaker failure timer setting has been assumed to be 90 ms for the 2-cycle
breaker 4. This could vary from one application to another. For a 3-cycle oil
breaker, a typical value is 150 ms.
Communication channel time depends on channel medium used.
Transient Stability
Chap.
The performance of protective relaying schemes during swings (electromechanical oscillations) and out-of-step conditions may be illustrated by considering
the simple two-machine system of Figure 13.36. The machines are assumed to be
represented by voltages of constant magnitudes behind their transient impedances (2
A
and 2'). The effect of swings on relaying quantities is analyzed by considering the
voltage, current, and apparent impedance measured by a relay at bus C (terminal of
machine A).
C
~ e n - ~ o +1-
Line
D
I
0Gen-B
EA and EB are voltages behind the transient impedances of the two machines.
EB is assumed to be the reference phasor, and 6 represents the angle by which EA
leads EB.
The current I is given by
Sec. 13.5
During a swing, the angle 6 changes. Figure 13.37 shows the locus of 2, as a function
of 6 on an R-X diagram, when EA=EB.
Electrical centre
'
j C Z c when 6=90
%?-&Lc~
_---
*--
Note:
-------
Zc when 8=60
91 6
Transient Stability
Chap. 1
When the voltage magnitudes EA and EB are equal, the locus of Zc is seen to
be a straight line which is the perpendicular bisector of the total system impedance
between A and B, i.e., of the impedance ZP The angle formed by lines from A and
B to any point on the locus is equal to the corresponding angle 6.
When 6=0, the current I is zero and Zc is infinite. When 6=180, the voltage
at the electrical centre (middle of total system impedance) is zero. Therefore, the
relay at C in effect will see a three-phase fault at the electrical centre. The electrical
centre and impedance centre coincide in this case.
If EA is not equal to EB, the apparent impedance loci are circles with their
centres on extensions of the impedance line AB. When EA>EB,the electrical centre
will be above the impedance centre; when EA<EB,the electrical centre will be below
the impedance centre. Figure 13.38 illustrates the shape of the apparent impedance
loci for three different values of the ratio EA/EB.
See. 13.5
91 7
After a system disturbance, the rates of change of the angle, voltage, and
current are quite variable. At first the rotor angle changes quickly, slowing down as
the peak angle is reached for a stable case; then the angle decreases and oscillates
with smaller changes until equilibrium is reached. If the system is unstable, the angle
increases gradually until it reaches 180, when a pole is slipped. Unless the system
is separated by protective systems, this is followed by repeated pole slips in rapid
The voltages and apparent impedances at points near the electrical centre
oscillate rapidly during pole slipping.
Actual loci of apparent impedances measured by distance relays are more
than for the idealized case we have considered; they depend on variations
in internal machine voltages, voltage regulator and speed governor actions, and
interactions between all the machines in the system as influenced by the
interconnecting network. Such loci can be readily determined for any given situation
by using a transient stability program. However, analysis of idealized cases involving
simple-system configurations is helpful in understanding the results obtained with
complex simulations.
Generalized impedance diagram
The concept of using the impedance diagrams for the analysis of distance relay
performance was initially developed by C.R. Mason and J.H. Neher [25,26]. They
considered the "swing line" on the R-X diagram for the case EA=EBand analyzed the
performance of different relays by plotting their characteristics on the same R-X
diagram. This was indeed a pioneering step in understanding the performance of
protective systems during power swings.
This concept was subsequently extended by other workers such as Edith Clarke
[27] by developing impedance loci for various values of the ratio EA/EB.They showed
that the impedance locus was a circle for each value of the ratio EA/EB,centred on the
system impedance line with radii and offset determined by the values of the voltage
ratio as shown in Figure 13.39. The specific case of EA/EB=I is the limiting case with
infinite radius and offset.
In reference 27, Edith Clarke also developed another family of curves
representing apparent impedance loci for constant angular separation (6). If the angle
between EA and EB is held constant while the voltage-magnitude ratio is varied, the
apparent impedance will trace a portion of a circle which passes through A and B, and
whose centre lies on the perpendicular bisector of the system-impedance line. These
characteristics are shown in Figure 13.40. Line AB is seen to be a portion of the circle
(with infinite radius) which represents O0 and 180" angular separation. It also
separates the right and left portions of each constant-6 circle into two characteristics
such that the separation angle for the parts differs by 180". For example, the 90"
circle on the right becomes the 270" circle on the left.
The two families of characteristics shown in Figures 13.39 and 13.40 can be
combined to form the generalized per-unit impedance diagram [24,27]. These
characteristics are helpful in visualizing and analyzing how the apparent impedance
varies in the general case when both the angular separation and the ratio of source
voltages are varied.
Transient Stability
1"
Sec. 13.5
919
Figure 13.41
I
Figure 13.42
The performance of protective relaying, using a timed second zone and not
using communication between line terminals, does not depend on relaying at the
opposite end of the circuit. Tripping occurs if the impedance locus at either end
remains in the second-zone characteristic sufficiently long to allow timed operation
or if the first-zone characteristic is entered.
Performance of schemes using pilot relaying will depend on impedances
measured at both ends.
Transient Stability
Chap,
(b)
' ~ a u l tarea
Sec. 13.5
Figure 13.44
The out-of-step tripping relays should not operate for stable swings. They must
detect all unstable swings and must be set so that normal load conditions are not
picked up.
922
Transient Stability
Chap. I
The majority (60 to 80%) of transmission line faults are of a transitory nature.
An example of such a fault is an insulator flashover due to high transient voltages
induced by lightning. After the line is de-energized long enough for the fault source
to pass and the fault arc to de-ionize, the line may be reconnected. Therefore,
common practice is to reclose the circuit breakers automatically to improve service
continuity.
Reclosing may be either single-shot (one attempt) or multishot (several
attempts) with time delay between each attempt. If the fault persists after two or three
attempts, the operator may attempt to reclose manually after some delay. If this fails,
it is indicative of a permanent fault and the line is taken out of service for repair. The
first attempt to reclose may be either high-speed or with time delay.
High-speed reclosure refers to the closing of circuit breakers after a time just
long enough to permit fault-arc de-ionization. The reclosure can be completed in less
than 1 second. The following are the benefits of high-speed reclosure:
e
Sec. 13.5
923
Minimization of the effect of line outage on critical load areas and restoration
of normal voltages.
However, high-speed reclosure may not be acceptable in all cases. Reclosure into a
permanent fault may cause system instability. Many utilities operate on the premise
that all circuit faults are permanent and therefore the system must successfully
withstand reclosing into a fault. In such cases, high-speed reclosure does not
contribute to increased flexibility of system operation.
In addition, high-speed reclosure of lines close to steam turbine generating
units may increase shaft fatigue. This is discussed further in Chapter 15.
In view of the above considerations, timed and supervised reclosure is often
used on the bulk transmission system. This ensures that specific conditions are
satisfied and a preferred sequence of reclosing of circuit breakers is followed. The
general procedure is as follows:
Potential is applied by closing one pre-selected breaker at one terminal after
a time delay of 5 to 10 seconds.
At terminals where out-of-synchronism conditions are a significant probability
(e.g., at points in the system with few lines), the other circuit breakers
associated with the line are automatically closed after the line has successfully
remained on potential with acceptable angles across open breakers (measured
by synchrocheck relays) for a period of at least 1 second.
Where the probability of an out-of-synchronism condition existing between
terminals is minimal, the circuit is automatically placed back on load by
closing all remaining breakers associated with the circuit after it has
successfully remained on potential for about 0.5 seconds.
13.5.7 Generator Out-of-Step Protection
For situations where the electrical centre is out in the transmission system, the
detection of an out-of-step condition and the isolation of unstable generator(s) are
accomplished by line protection.
However, for situations where the electrical centre is within the generator or
step-up transformer, a special relay must be provided at the generator. Such a situation
occurs when a generator pulls out of synchronism in a system with strong
transmission. A low excitation level on the generator (EA<EB)also tends to contribute
to such a condition.
924
Transient Stability
Chap, 13
of slip of its poles. The high amplitude currents and off-nominal frequency operation
could result in winding stresses and pulsating torques that can excite potentially
damaging mechanical vibrations. There is also a risk of losing the auxiliaries of the
.
affected unit as well as the auxiliaries of nearby stable units.
In order to avoid these adverse effects on the unit and the rest of the system
it is desirable to have an out-of-step relay that will trip the unit; the tripping should
be restricted to disconnection of the machine from the system by tripping the
generator breaker, rather than initiating a complete shutdown of the unit. This allows
resynchronization of the generator as soon as conditions stabilize.
.
Sec. 13.5
Infinite
B/
/
t$ 5 bus
2s
--C
zaPP
-
locus
.(
YA
Transient Stability
Chap. 13
Blinder element
Example 13.6
For the following unstable case simulated in Example 13.2, plot the apparent
impedance measured by a distance relay located at the HT bus:
-
Identify the location of the electrical centre and compute the effective machine
reactance.
Solution
Figure E13.22 shows the locus of apparent impedance computed during the timedomain simulation.
From the apparent impedance locus plot, we see that the electrical centre (C) lies
between the HT bus (H) and the infinite bus (B). The reactance between points H and
C, as measured on the plot, is equal to 0.055 pu. Therefore,
Sec. 13.5
Solving, we get
0.3 0.2 -
./'(prefault)
r t=1.0 s
Electrical
centre
0-1 - t=1.71 s
Equivalent circuit
I .b R in Pu
0.0
X,=O. 15
-0.2 -0.3 -0.4 -
AXJ
928
Transient Stability
Chap, 1
-2-o
Sec. 13.5
929
When excitation is lost, the field flux linkage and the effective internal voltage
gadually decrease depending on the field circuit time constant. This results in a
gadual reduction in active power output, accompanied by an increase in rotor angle.
In addition, the machine absorbs increasing amounts of reactive power as the field
flux decays. Referring to the two-machine equivalent circuit of Figure 13.36, what we
have is a situation where the internal voltage EA is decreasing and the rotor angle 6
is increasing. The loss-of-excitation characteristic is, therefore, some combination of
the two families of curves shown in Figures 13.39 and 13.40. The actual path traced
depends on the relative rates of variation in EA and 6, which in turn depend on the
initial generator output, transmission strength, and field circuit time constant. The
overall effect on the apparent impedance at the machine terminals is as depicted in
Figure 13.48.
With a high initial output, the rotor angle advances more rapidly and the
endpoint shown in the figure is reached more quickly, typically within about 10 s. The
slip is also relatively high, and therefore the impedance locus terminates at a value
corresponding to the average of subtransient d- and q-axis reactances.
On the other hand, with a low initial output, the rotor angle advances slowly
and the slip at the endpoint is small. The impedance locus terminates at a value
corresponding to the average of d- and q-axis synchronous reactances.
The locus of the endpoints of the loss-of-excitation characteristic is shown in
Figure 13.49. It ranges between the subtransient reactances and the synchronous
reactances. Although these extreme values are not actually reached, they are helpful
in identifying the LOE relay characteristic. By encompassing the locus of the
endpoints as shown in Figure 13.49, the relay will operate before the generator first
starts to slip poles.
-R
Nearly equal to X:
Locus of endpoints
of LOE characteristics
ly equal to Xd
Relay characteristic
Figure 13.49 LOE relay and system characteristics
930
Transient Stability
Chap. 13
Loss-of-excitation relays
The characteristics of an LOE relay proposed by Mason in reference 29 are
shown in Figure 13.49. It is an offset mho relay. The offset is equal to one-half ofthe
transient reactance of the generator. The total reach of the relay is equal to the
synchronous reactance Xd.
The offsetting of the relay characteristic from the origin provides selectivity
against improper operation during power swings or out-of-step conditions. In
extreme case of zero external impedance, the electrical centre (intersection of the
swing locus with the system impedance locus) is at the centre of the generator
impedance Xj.Therefore, the LOE relay characteristic is offset by XJ12.
The offset mho relay with the above characteristic has been widely used and
has generally performed satisfactorily.
To provide greater selectivity against false operation during underexcited
operation, stable transient swings, or system disturbances causing underfrequency
operation, the use of two offset rnho relays with characteristics as shown in Figure
13.50 is proposed by J. Berdy in reference 30.
Reference 3 1 describes a more discriminating relay developed at Ontario
Hydro in 1978. This relay can prevent unit trip for recoverable excitation failures and
trip the unit more rapidly for unrecoverable failures. The relay uses generator field
and terminal voltages as inputs. It operates on the general principle that when the
terminal voltage is high the field voltage should be low and, conversely, when the
terminal voltage is low the field voltage should be high. For a detected excitation
problem, the relay will initially try to transfer voltage control to an alternate
AVRIexciter. If this does not rectify the problem, the relay trips the unit. Figure 13.51
shows a block diagram of the relay. The relay also includes selective overexcitation
protection as shown by the dashed blocks in the figure.
,,
-xl
Figure 13.50 LOE protection with two mho relays
Sec. 13.5
I Overvoltage I---;
i I Timer /-----lANd1-,
I
I------
!--1
detector
Undervoltage - Timer
detector
voltage
,----__-------__-
Relay
driver
II
(Excitation
regulator
transfer)
1
I;
AC
generator
terminal
voltage
DC field
.--I
1
r
'
Undervoltage
detector
Relay
Timer driver
-G
Timed trip
(Generator
trip)
Example 13.7
In this example, we examine the response of the system of Example 13.2 to loss of
generator excitation.
(a)
If the field of the equivalent generator representing the four units of the plant
is suddenly shorted (directly, with zero external resistance), determine the
locus of apparent impedance measured by an LOE relay located at the
generator terminals. Assume that the LOE relay has a characteristic similar to
that shown in Figure 13.49. Show the corresponding time responses of
generator terminal voltage, active and reactive power outputs, and rotor angle.
(b)
Determine the locus of apparent impedance if the field of only one of the four
generators in the plant is shorted (directly). The other generators operate
normally with excitation control.
Assume that in each of the above cases the initial system operating condition is the
same as in Example 13-2.
Solution
(a) The apparent impedance locus computed by using a transient stability program
is shown in Figure E13.23. The generator is represented in detail including
amortisseur circuits, and shorting of the field is simulated by setting the field voltage
E/, to zero at t=1.0 s.
Transient Stability
Chap. 13
The corresponding time response plots of Et, P, Q and 6 are shown in Figure E13.24.
These plots are helpful in explaining the shape of the apparent impedance locus, if we
note that
L
where
From Figure E13.23, we see that the apparent impedance locus enters the LOE
characteristic at t=3.4 s, i.e., 2.4 s after the field is shorted. As we see in Figure
E13.24(b), this corresponds to the time when the generator 'has slipped one complete
cycle. As the generator continues to slip poles, P, Q, and E, oscillate in rapid
succession.
(b) Figure E13.25 shows the apparent impedance locus when the field of one of the
generators is shorted. The other three generators continue to operate normally with
AVR and PSS controlling their excitation.
13.5
Time in seconds
(a) Time response' of terminal voltage E,
Time in seconds
(c) Time response of active power P
Time in seconds
(b) Time response of rotor angle 6
Time in seconds
(d) Time response of reactive power Q
Transient Stability
Chap. 13
The system
The area of interest for the system under study is shown in Figure 13.52. It
consists of an 8 unit 7,000 MW nuclear plant connected to the rest of the system
through two 500 kV double circuit lines and three 230 kV double circuit lines. The
system model, which represents a portion of a large intercohnected power system,
,- consists of
o
e
o
Sec. 13.6
Transient Stability
936
Chap, 1
The contingency
Event
No disturbance
100
164
5000
Terminate simulation
Simulation
Sec. 13.6
501
-50.
0.0
0.5
1.O
/Remote
1.5
generator
2.0
2.5
937
3.0
Time in seconds
Figure 13.53 Rotor angle time response
becomes generation deficient, and hence the absolute angles of all machines in the
system drift slightly. In this regard, using relative angles rather than absolute angles
is ofien a better choice since it permits one to easily observe the relative motion of
rotors between machines. The LLG fault results in the loss of one 500 kV doublecircuit line, leaving the entire 7,000 MW plant connected to the system through a
significantly weaker (i.e., higher impedance) path consisting of one double-circuit 500
kV line and three 230 kV double-circuit lines. The plant is unable to transfer full
output through this path, and the resulting accelerating power leads to instability.
Transient Stability
-0.1
-0.15
-0.1
-0.05
0.0
0.05
0.1
Chap, 1
0.15
Resistance R in pu
Figure 13.54 Unit G3 out-of-step protection
Figure 13.55 and Figure 13.56 show the line protection performance for relays
on circuit 3 at buses 1 and 7, respectively. The apparent impedances measured at each
end looking toward the opposite ends are shown on the plots. The relays are mho
distance relays and have zone 1 coverage of about 75% of line length and zone 2
overreach of about 125% of line length. These plots show that the apparent impedance
enters the zone 2 relays at bus 1 and enters the zone 1 and zone 2 relays at bus 7. The
zone 1 relay at bus 7 would trip circuit 3 at bus 7 and send a transfer trip signal to
the breakers at bus 1 which would then trip circuit 3 at bus 1. This logic also holds
true for the companion 500 kV circuit (#4) which would be tripped in an identical
manner. Following the loss of the 500 kV circuits (at approximately 0.8 s), the
remaining 230 kV circuits would become extremely overloaded and would be lost
through protection actions, thereby completely isolating the unstable plant from the
system.
From Figure 13.55, we can determine that the impedance swing crosses the
circuit at a point about 84% of the line length from bus 1. This point represents the
electrical centre following the disturbance and is theoretically where separation occurs
(although the entire circuit is tripped). Figure 13.57 shows, plots of bus voltage
measured at buses 1 and 7, and at the electrical centre. The voltage at the electrical
centre leads the other voltages in collapsing as instability occurs.
13.6
Resistance R in pu
Figure 13.55 Line protection (circuit 3) at bus 1
Relay characteristic
Resistance R in pu
Figure 13.56 Line protection (circuit 3) at bus 7
Transient Stability
Chap. 13
----------
1.0 0.8
0.6
0.4
\
'
+
'
-.--_
--___
++
+$,
Bus 7
,/'
2%.
Electrical centre
\'.
-.
---_
Y,
\'\
\',
Y-,
+.
\'
0.2
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
\\
\ \'
\\
0.8
Time in seconds
Figure 13.57
Reduction of the pre-contingency output of the plant. If one unit is taken off
line (with output redispatched to other units in the external system), the total
plant output will be reduced sufficiently to ensure stability following the LLG
fault. Curve 2 in Figure 13.58 shows the absolute rotor angle for unit G3 for
this case. It shows that the plant remains in synchronism. This solution to
instability is, however, generally costly because available energy is kept idle
or bottled in the plant.
(b)
Sec. 13.7
250
/t 8 units, no GR
V3
a,
2
gJ 200 -
i
:
(curve 1)
+-,
41 100-
\
\
Q)
\'
'
/'
'
0**
+-,
--__0'
50-
GR
-_
7-----
8 units, 2 with GR
(curve 3)
generation rejection
Time in seconds
Figure 13.58 Unit G3 rotor angle responses with
and without generation rejection
942
Trans.ient Stability
Chap. 13
rim of the bowl is irregular in shape so that different points on the rim have different
heights.
SEP
w
Initially the ball is resting at the bottom of the bowl, and this state is referred
to as the stable equilibrium point (SEP). When some kinetic energy is injected into
the ball, causing it to move in a particular direction, the ball will roll up the inside
surface of the bowl along a path determined by the direction of initial motion. The
point where the ball will stop is governed by the amount of kinetic energy initially
injected. If the ball converts all its kinetic energy into potential energy before reaching
the rim, then it will roll back and eventually settle down at the stable equilibrium
point again. However, if the kinetic energy injected is high enough to cause the ball
to go over the rim, then the ball will enter the region of instability and will not return
to the stable equilibrium point. The surface inside the bowl represents the potential
energy surface, and the rim of the bowl represents the potential energy boundary
surface (PEBS).
Two quantities are required to determine if the ,ball will enter the instability
region: (a) the initial kinetic energy injected and (b) the height of the rim at the
crossing point. The location of the crossing point depends on the direction of the
initial motion.
Sec. 13.7
943
same manner as the ball rolling up the potential energy surface. To avoid instability,
the system must be capable of absorbing the kinetic energy at a time when the forces
on the generators tend to bring them toward new equilibrium positions. This depends
on the potential energy-absorbing capability of the postdisturbance system. For a
given postdisturbance network configuration, there is a maximum or critical amount
of transient energy that the system can absorb. Consequently, assessment of transient
stability requires
(a)
(b)
For a two-machine system the critical energy is uniquely defined, and the TEF
analysis is equivalent to the equal-area criterion described in Section 13.1. This is
illustrated in Figure 13.60, which shows two plots, both having rotor angle (6) as the
ordinate [40]. The upper plot illustrates the equal-area criterion in which the critical
clearing angle (63 is established by equality of areas A l and A2. The lower plot
illustrates the transient energy method which can be used to specify the critical
clearing angle in terms of potential and kinetic energy. The kinetic energy gained
during the fault-on period is added to the potential energy at the corresponding rotor
angle, and the sum is compared to the critical potential energy to determine stability.
Given a disturbance, there is a stable equilibrium point for the postfault
system. A region of attraction can be defined for this postfault SEP as shown in
Figure 13.61. Any postfault system trajectory with the state of the system at fault
clearing (xcl) inside this region of attraction will eventually converge to the SEP, and
the system is said to be stable. On the other hand, if xcl lies outside the region of
attraction, the postfault system will not converge to the stable equilibrium point, and
the system is said to be unstable.
The state of the system at fault clearing (x,J can be described by the value of
the energy function evaluated at xcl, i.e., V(xcl). Hence the direct method solves the
stability problem by comparing V(xcl) to the critical energy V., The system is stable
if V(xcl) is less than V, and the quantity V,-V(xcl) is a good measure of system
relative stability. This quantity is defined as the transient energy margin.
The quantity V(xcl) measures the amount of transient energy injected into the
system by the fault while the critical energy measures the strength of the postfault
system. More precisely, the critical energy measures the energy-absorbing capability
of the postfault system.
Referring to Figure 13.61, if the rotor oscillates within the range 6,, to 6,,, the
system will remain transiently stable. If the rotor swings beyond this range, the system
will become unstable. Hence the two points 6,1 and 6,2 on the potential energy curve
form a boundary to all stable rotor angle trajectories. This boundary is called the
potential energy boundary surface (PEBS) and the points on this boundary are local
potential energy peaks.
Transient Stability
Infinite
bus
t
(a) System configuration
Power 1
Energy
Critical energy
PE(6,)
Chap. 13
Sec. 13.7
946
Transient Stability
Chap. 13
have been several attempts to extend the method to include more detailed models
[46,47]. This, however, is still at a developmental stage and f~rtherresearch efforts
are required to improve the accuracy and reliability of the method with detailed
models. Here, we will limit our discussion to the simplified system models.
With the generator transient reactances and load admittances included in the
node admittance matrix, we may write
where YR is the reduced admittance matrix with all nodes other than the generator
internal nodes eliminated, EG is the generator internal source voltage vector, and rG
is the generator current vector.
Let the internal voltage of the ith generator in phasor notation be given by
For a system with n machines, the active output of the ith generator is given by
For the application of the TEF method, it is convenient to describe the transient
behaviour of the system with the generator rotor angles expressed with respect to the
inertial centre of all the generators. The position of the centre of inertia (COI) is
defined as
where HT is the sum of the inertia constants of all n generators in the system.
The motion of the COI is determined by
Sec. 13.7
and
with
where
P,,
D:.
= E.E.G~
I/
1 I
oo = synchronous speed in elec. radls
Amco,
The motion of the generators with respect to the COI can be expressed by defining
=
6,-GcOI
rad
(13.93)
and
The equations of motion of the ithmachine in the COI reference frame are
where oi is the pu speed of the ith machine with respect to the COI.
The energy function V describing the total system transient energy for the
postdisturbance system is defined as
Transient Stability
Chap. 13
where
8:
ith generator
ZPm~(0, - 0:) :
(c)
xx cv(cosOv- cos0;)
I .
(d) x x ~ ~ c o s 0 ~+ej)
d (:0 change
~
in dissipated energy of all branches
The first term is called the kinetic energy (Vke)and is a function of only generator
speeds. The sum of terms 2, 3, and 4 is called the potential energy (Vpe) and is a
function of only generator angles.
The transient stability assessment involves the following steps:
1.
2.
3.
Time-domain simulation is run up to the instant of fault clearing to obtain the angles
and speeds of all the generators. These are used to calculate the total system energy
(VcJ at fault clearing.
The calculation of V, i.e., the boundary of the region of stability, is the most
difficult step in applying the TEF method. Three different approaches are briefly
described below.
Sec. 13.7
949
Transient Stability
Chap, 1
There are many solutions to the above equations, one of which is the SEP. All other
solutions are UEPs. The controlling UEP is calculated by solving the following
minimization problem:
min
F=C&
i : l }
An approximate UEP based on the MOD is used as a starting point for the solution,
Numerical problems are sometimes encountered when one is solving for the
controlling UEP based on the above approach. A robust procedure based on the
potential energy boundary surface crossing point suggested in reference 44 is
described below.
(c) The boundary of stability-region-based controlling UEP (BCU) method
Earlier CUEP methods face serious convergence problems when solving for
the controlling UEP, especially when the system is highly stressed or highly
unstressed, or the mode of system instability is complex. These problems usually arise
if the starting point for the CUEP solution is not sufficiently close to the exact CUEP.
Some of the convergence problems can be overcome by the BCU method which has
the capability of producing a much better starting point for the CUEP solution.
In reference 44, the boundary of the stability region is defined as the union of
the stable manifolds of the UEPs on the boundary. Any trajectory starting from a
point on the boundary will converge to one of the UEPs as time increases. By making
use of this property, the BCU method computes the controlling UEP of a power
system through computing the CUEP of its associated reduced system.
The dynamic behaviour of a multimachine power system represented by
classical generator models and constant impedance loads is described by Equations
13.95 and 13.96. The associated reduced system model for this multimachine power
system is defined as
The equilibrium points of the reduced system are the same as those of the original
system. The BCU method computes the CUEP as follows:
Sec. 13.7
951
(i)
(ii)
Starting with the exit point, the reduced system is simulated in time domain
until the quantity C lgi 1 reaches its first local minimum. The rotor angles at
this point constitute the starting point for the CUEP solution.
(iii)
The controlling UEP is computed using any robust numerical technique. The
UEP so obtained is the controlling UEP relative to the fault-on trajectory.
(4 Sustained-fault approach
[42]
The calculation of the controlling UEP for large systems is very timeconsuming. Under severe system conditions, the solution process may either fail to
converge or converge to the wrong UEP. The sustained-fault approach was developed
to avoid the calculation of the controlling UEP.
The critical energy is determined as follows:
1.
A time domain simulation with a sustained fault is run until the system crosses
the potential energy boundary surface (PEBS). The crossing of the PEBS is
indicated by the potential energy reaching its maximum.
2.
The potential energy at the crossing of the PEBS is taken as the critical energy
for that particular fault location.
Example 13.8
In this example, we illustrate the use of the TEF method for transient stability analysis
(b)
(c)
Calculate the postdisturbance system SEP, UEP, and the critical energy
V'.
Transient Stability
(d)
Chap. 13
Calculate the energy at fault clearing with tc=0.07 s, 0.086 s, and 0.087 s.
Determine the system stability for each of the three f3ult durations.
Solution
(a) The postdisturbance reduced equivalent circuit of the system is shown in Figure
E13.26.
and A oCol
= 0 . Consequently, el = 6
Since GI2=0,D12=O and P A =P,, =0.9. The postdisturbance dynamic equations of the
system, corresponding to Equations 13.95 and 13.96, are
Hence,
0;
0.9552 rad
54.73'
For this simple system there is only one postdisturbance UEP, which is given by
13.7
n-8;
2.1864 rad
953
125.27"
The critical energy, which equals the system potential energy at the postdisturbance
UEP, is
v,
0.1651
(d) To compute the energy at fault clearing, we will use the generator angles
computed in Example 3.1.
(i) With tc=0.07 s, 0;=48.58"=0.8479 rad, and
is
stable; the
stability
margin
is
Vcr-Vc,=
Transient Stability
Chap. 13
REFERENCES
[I]
W.D. Stevenson, Elements of Power System Analysis, 3rd Edition, McGrawHill, 1975.
[2]
R.T. Byerly and E.W. Kimbark, Stability of Large Electric Power Systems,
IEEE Press, 1974.
[3]
E.R. Laithwaite and L.L. Freris, Electric Energy: Its Generation, Transmission
and Use, McGraw-Hill (UK), 1980.
[4]
A. Ralston and H.S. Wilf, Mathematical Methods for Digital Computers, John
Wiley & Sons, 1962.
[5]
B. Carnahan, H.A. Luther, and J.O. Wilkes, Applied Numerical Methods, John
Wiley & Sons, 1969.
[6]
References
955
[7]
[a]
[9]
[lo]
11
[12]
H.W. Dornmel and N. Sato, "Fast Transient Stability Solutions," IEEE Trans.,
Vol. PAS-91, pp. 1643-1650, JulyIAugust 1972.
[13]
[14]
[15]
[16]
Edith Clarke, Circuit Analysis of AC Power Systems, Vol. I, John Wiley &
Sons, 1943.
1171
P.M. Anderson, Analysis of Faulted Power Systems, Iowa State Press, Ames,
Iowa, 1973.
[18]
[19]
[20]
Transient Stability
Chap. 13
C.R. Mason, The Art and Science of Protective Relaying, John Wiley & Sons,
1956.
General Electric Company, Use of the R-X Diagram in Relay Work,
Philadelphia, Pa., 1966.
C.R. Mason, "Relay Operation during System Oscillations," AIEE Trans., Val.
56, pp. 823-832, 1937. Discussion by J.H. Neher, pp. 1513-1514. Closure by
C.R. Mason, Vol. 57, pp. 111-114, 1938.
J.H. Neher, "A Comprehensive Method of Determining the Performance of
Distance Relays," AIEE Trans., Vol. 56, pp. 833-844, 1937.
Edith Clarke, "Impedances Seen by Relays during Power Swings with and
without Faults," AIEE Trans., Vol. 64, pp. 372-384, 1945. Discussion by A.J.
Mccomell, p. 472, June Supplement 1945.
IEEE Working Group Report, "Out of Step Relaying for Generators," IEEE
Trans., Vol. PAS-96, pp. 1556-1564, SeptemberIOctober 1977.
C.R. Mason, "A New Loss-of-Excitation Relay for Synchronous Generators,"
AIEE Trans., Vol. 68, pp. 1240-1245, 1949.
J. Berdy, "Loss of Excitation Protection for Modern Synchronous Generators,"
IEEE Trans., Vol. PAS-94, pp. 1457-1463, SepternberlOctober 1975.
D.C. Lee, P. Kundur, and R.D, Brown, "A High Speed, Discriminating
Generator Loss of Excitation Protection," IEEE Trans., Vol. PAS-98, pp.
1895- 1898, November/December 1979.
P.C. Magnusson, "Transient Energy Method of Calculating Stability," AIEE
Trans., Vol. 66, pp. 747-755, 1947.
P.D. Aylett, "The Energy Integral-Criterion of Transient Stability Limits of
Power Systems," Proc. IEE, Vol. 105c, No. 8, pp. 527-536, September 1958.
G.E. Gless, "Direct Method of Lyapunov Applied to Transient Power System
Stability," IEEE Trans., Vol. PAS-85, No. 2, pp. 159-168, February 1966.
References
957
Transient Stability
Chap. 1
A.A. Fouad, V. Vittal, Y-X. Ni, H.M. Zein-Eldin, E. Vaahedi, H.R. Pota, K.
Nodehi, and J. Kim, "Direct Transient Stability Assessment with Excitation
Control," IEEE Trans., Vol. PWRS-4, No. 1, pp. 75-82, February 1989.
V. Vittal, N. Bhatia, A.A. Fouad, G.A. Maria, and H.M. Z e i n - ~ l d i ~ ,
"Incorporation of Nonlinear Load Models in the Transient Energy Function
Method," IEEE Trans., Vol. PWRS-4, No. 3, pp. 1031-1036, August 1989.
T. Athay, R. Podmore, and S. Virmani, "A Practical Method for the Direct
Analysis of Transient Stability," IEEE Trans., Val. PAS-98, No. 2, pp. 573584, MarchIApril 1979.
A.A. Fouad, V. Vittal, and Taekyoo Oh, "Critical Energy for Transient
Stability Assessment of a Multimachine Power System," IEEE Trans., Val.
PAS- 103, pp. 2 199-2206, 1984.
C.K. Tang, C.E. Graham, M. El-Kady, and R.T.H. Alden, "Transient Stability
Index from Conventional Time Domain Simulation," Paper 93SM487-9PWRS,
presented at the 1993 IEEE PES Summer Meeting, Vancouver, July 1993.