Download A Method for Real Time Voltage Stability Monitoring in Sub

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Tube sound wikipedia , lookup

Wireless power transfer wikipedia , lookup

Decibel wikipedia , lookup

Immunity-aware programming wikipedia , lookup

Mechanical-electrical analogies wikipedia , lookup

Power over Ethernet wikipedia , lookup

Power inverter wikipedia , lookup

Rectifier wikipedia , lookup

Scattering parameters wikipedia , lookup

Power factor wikipedia , lookup

Audio power wikipedia , lookup

Current source wikipedia , lookup

Electric power system wikipedia , lookup

Two-port network wikipedia , lookup

Surge protector wikipedia , lookup

Pulse-width modulation wikipedia , lookup

Power MOSFET wikipedia , lookup

Stray voltage wikipedia , lookup

Metadyne wikipedia , lookup

Variable-frequency drive wikipedia , lookup

Electrification wikipedia , lookup

Electrical substation wikipedia , lookup

Power electronics wikipedia , lookup

Amtrak's 25 Hz traction power system wikipedia , lookup

Switched-mode power supply wikipedia , lookup

Three-phase electric power wikipedia , lookup

Voltage optimisation wikipedia , lookup

Buck converter wikipedia , lookup

History of electric power transmission wikipedia , lookup

Islanding wikipedia , lookup

Power engineering wikipedia , lookup

Mains electricity wikipedia , lookup

Alternating current wikipedia , lookup

Nominal impedance wikipedia , lookup

Zobel network wikipedia , lookup

Impedance matching wikipedia , lookup

Transcript
A Method for Real Time Voltage Stability
Monitoring in Sub-transmission Networks
D. T. Duong and K. Uhlen
S. Løvlund
Department of Electric Power Engineering
Norwegian University of Science and Technology
Trondheim, Norway
Statnett
Alta, Norway
Abstract—This paper presents a real-time voltage
stability monitoring method based on estimation of
maximum power transfer. The approach solely requires
information from PMUs installed at concerned load buses
at substations in addition to available information of
system topology and load power obtained from
SCADA/EMS or state estimator. The proposed algorithm
estimates the Thevenin impedance seen at the load bus and
consequently the maximum loadability. This power
transfer limit is used as an indicator for voltage stability
assessment when compared to the load power. It is
emphasized that only topology of the considered subsystem is needed, not the whole network. This feature
makes the method viable for online implementation since
the computation requirement is insignificant. In principle,
the proposed methodology is not limited to specific grids in
terms of voltage level, but it is especially suitable for subtransmission networks, where loads are supplied by bulk
power source and voltage instability is imposed by long
transmission lines. The proposed approach is validated by
simulations on the Norwegian transmission grid.
Keywords—Admittance matrix, maximum loadability, PMU,
Thevenin impedance, voltage stability.
I.
INTRODUCTION
The development towards increasing use of renewable
energy sources and smart grid applications represent a
paradigm shift in power system operation. The transmission
system must be more flexible and well monitored to adapt to
intermittent generation and dynamic load pattern. This scenario
requires new real time monitoring tools to timely identify
operation limits. For those systems with long and heavily
loaded transmission lines, voltage stability is one of the factors
that set limits on power transfer. There have been many works
on voltage stability, but most of them are suitable for off-line
applications. The method of continuation power flow [1]
requires repetition of load flow calculation of the whole system
to compute the point of maximum power transfer. This task is
time-consuming, especially in large power system. In addition,
the method does not include dynamic system behaviors. The
other approach proposed by [2] is to analyze the sensitivity of
the Jacobian matrix. The study figures out that as the system is
approaching the “nose point”, the eigenvalue of this Jacobian
matrix is going to zero. Hence, eigenvalues of this matrix can
be used to evaluate voltage stability condition. In order to
implement this method for online application, we need fast and
full observer of the whole system, which is not obtainable at
present.
Phasor measurement units (PMUs) open up opportunities
for new research in online applications in monitoring, control
and wide area protection. Regarding online voltage stability
assessment, [3] introduces a novel approach based on the
Thevenin impedance and the impedance matching criterion.
The method uses phasor measurements at the load bus to
estimate the Thevenin impedance seen at that one. Comparing
the load impedance with the Thevenin impedance gives the
indication of voltage instability. A further development in [4]
finds the Thevenin impedance with repetitive process of
estimation and correction until the algorithm converges.
Reference [5] evaluates voltage stability, based on the whole
system topology and information of all load power, by forming
“single-port circuit” for each concerned load. From this circuit,
a new impedance is computed. Its function is similar to that of
the Thevenin impedance in terms of assessing voltage stability.
This paper proposes another method to find the Thevenin
impedance seen at load buses. Similar to [5], the approach
requires system topology and load power. However, it needs
only information of the concerned sub-system. The study also
examines the case, in which estimation of equivalent
impedance based on the Thevenin theorem fails to evaluate the
maximum loadability.
The rest of this paper is organized as follows. Section II
reviews voltage stability assessment based on the Thevenin
impedance, followed by the proposed method in Section III.
The next section describes a method for system reduction and
identification of the boundary of the concerned sub-system.
Section V examines the effect of active power limitation of
generators. Section VI shows simulation results, and finally
conclusions are drawn in Section VII.
II.
REVIEW OF VOLTAGE STABILITY MONITORING
BASED ON THE THEVENIN IMPEDANCE
Given a simple two-bus system as shown in Fig. 1, where
ETh is a voltage source behind ZTh, an impedance denoting a
transmission line. The load is considered as the impedance ZL,
which is determined by the ratio of the voltage ̃ to the current
̃ . Regardless of whether active or reactive power is extracted
from the grid at the load bus, it can be easily shown that the
maximum loadability is reached when the load impedance
magnitude is equal to that of the impedance ZTh,

|
|
|
|

VL
ZTh
transfer can be computed. Finally, the indicator for real time
voltage stability is established either by comparing the load
impedance to the Thevenin impedance by (2) or identifying
power margin by (5). Overall, the major challenge is to find the
Thevenin impedance, which is presented in the following
sections.
IL
ETh
ZL
III.
Figure 1. The Thevenin equivalent circuit.
Consequently, by comparing the two impedances, we can
establish the Impedance Stability Index (ISI) at the load bus as
[7]
ISI 

ZTh
ZL
.

Furthermore, when the ZTh is known, ETh can be identified
by

ETh  VL  I L ZTh , 

where VL and I L are voltage and current phasors of the load.
PROPOSED METHOD FOR ESTIMATION OF THE
THEVENIN EQUIVALENCE
The main idea of the proposed approach is to use available
information of system topology and load power, for example
obtained from SCADA or existing state estimator, to compute
the Thevenin impedance seen by the concerned load buses,
where PMUs are installed to measure phasors of load voltage
and currents. These measurements are utilized to facilitate real
time estimation of maximum loadability at each monitored bus.



Let us consider a simple grid in Fig. 2. The generator G1
connected to bus #1 through the step-up transformer T1 is
supplying the two loads L1 and L2 at bus #1 and #2
respectively; meanwhile Line 1 connects bus #1 and #2. Since
the focus is on voltage stability, it is reasonable to assume that
G1 has unlimited active power. From Fig. 2, the electric
equivalent circuit of this system can be depicted in Fig. 3,
where T1, Line 1, L1 and L2 are represented by impedances ZT,
ZL, ZL1, and ZL2 respectively. For the sake of illustration, shunt
admittances of Line 1 are neglected.
The maximum power transfer occurs at the instant when
(1) is fulfilled. Therefore, this maximum loadability point is
estimated by [5]
#1
T1
L2
G1
Smax 


ETh2  ZTh   imag  ZTh  sin   real  ZTh  cos  
, 
2
2 imag  ZTh  cos   real  ZTh  sin  


#2
Line 1
L1
Figure 2. A simple two-node grid.
ZT
ZL
#1
#2
where  is the load power angle.
After crossing the maximum loadability limit, the load
might come to voltage collapse if it contains constant power or
constant current load components. Obviously, different load
models exhibits various voltage collapse mechanisms,
depending on the specific load composition. Nonetheless,
operation beyond the limit is not expected. From voltage
stability monitoring point of view, it is reasonable to consider
the maximum power transfer as the limit for voltage instability.
Thus, the distance from the current operating point to voltage
instability is the power margin to the estimated maximum
loadability determined by [5]

S  SL
margin  max
100%, 
SL

where SL is the current apparent power of the load.
In a meshed power system, the main task of online voltage
stability monitoring method is to estimate properly the
Thevenin impedance of the rest of the system, which is one of
the main contributions of this article. Then, with local phasor
measurements of voltage and current, the maximum power
E1
Figure 3.
ETh
ZL1
ZL2
Equivalent circuit of the simple grid.
According to the Thevenin theorem, the equivalent
impedance of the system seen at bus #2 is
ZTh  Z L 
Z L1 ZT
Z L1  ZT
(6)
In addition, from Fig. 2, the admittance matrix of the grid is
(not including the load)
1
 1
Z Z
T
Y  L

1
 
ZL

1 
ZL 

1 

ZL 

(7)
Let us modify the diagonal elements of the matrix Y by
adding the load admittance to establish a new matrix
1
1
 1
Z Z Z
T
L1
Yeq   L

1


ZL

1 
ZL 

1 

ZL 

(8)
Define the impedance matrix Zeq as the inversion of Yeq,
Zeq  Yeq1.
(9)
The diagonal element in the second row and column of Zeq
is
Z eq  2, 2   Z L 
Z L1 ZT
,
Z L1  ZT
(10)
which is by definition equal to the Thevenin impedance seen at
bus #2 in (6) (in general, the diagonal element in ith row and ith
column of the impedance matrix is the Thevenin impedance
seen at the ith bus in the grid [6]). After the Thevenin
impedance is identified, the maximum loadability is computed
by (4).
It is noticed that in order to estimate the Thevenin
impedance seen by bus #2, the load impedance at bus #1 must
be integrated into the admittance matrix, but not at bus #2.
Then the modified admittance matrix is inverted. If bus #1 is
now a considered bus, the matrix inversion must be conducted
again, but load impedance at bus #2 is integrated this time.
Obviously, it is not an efficient method. Let us make a small
modification of the matrix Yeq in (8) by adding all load
admittances into the admittance matrix, resulting in
1
1
 1
Z Z Z
T
L1
Yeq   L

1


ZL



.
1
1 


ZL ZL2 

1
ZL
(11)
Inverting Yeq gives us the impedance matrix, whose
diagonal elements are impedances seen by the corresponding
buses. These impedances are indeed the actual Thevenin
impedance in parallel with the load impedance at each load
bus. Let us take bus #2 as an example. After inverting Yeq, the
equivalent circuit at this bus is shown in Fig. 4. Here Z'Th is the
impedance in the second row and column of the impedance
matrix. However, the impedance we intend to find is ZTh, not
Z'Th. In this case, ZTh can be found by
Z Z'
'
ZTh
 ZTh Z L 2  ZTh  L 2 Th' .
(12)
Z L 2  ZTh
#1
Z’Th
#2
E’Th
Figure 4. Thevenin equivalent circuit, including all load impedances.
In summary, the admittance matrix is modified by adding
all load admittances to the corresponding diagonal elements.
Then matrix inversion is implemented. For those concerned
buses, the Thevenin impedances are found by the modification
in (12). By doing this way, we need to invert the admittance
matrix only once.
IV.
SYSTEM SIMPLIFICATION
A. Determination of Sub-System Boudary
As described in Section III, the proposed method requires
the system topology of the whole system. It might be
impractical in a real power system and demand huge
computation effort, especially for online or near-online
applications. However, voltage stability is commonly
considered more of a local problem. Voltage collapse occurring
at a particular load bus is related to constraints of power lines,
transformers and generators that are feeding the load. Those
generators that are electrically far from the concerned loads
have very small or no impact on the study area. Obviously, it is
not necessary to include these generators. For a particular study
area that is prone to voltage instability, we can determine a
boundary that fulfills the condition that generators and
boundary nodes secure sufficient active power supply.
Consequently, voltage collapse in the study area is caused by
power lines, generators and loads in the concerned area; the
rest of the system can be neglected. The idea is illustrated in
Fig. 5. Assume that the study area is susceptible to voltage
collapse due to high load demand or loss of transmission lines.
When the sub-system is approaching voltage instability, bus B1
and B2 can be considered as the boundary nodes if voltage at
bus B1 and B2 are around nominal value since they are strong
buses. This allows us to simplify the study sub-system to the
small area as shown in Fig. 5. If bus B1 and B2 do not meet the
aforementioned assumption, the boundary should be expanded
to bus B3, B4 and B5. The boundary selection goes on until the
condition is met. This process can be done with off-line
analysis.
An example is shown in Fig. 6. The sub-system is supplied
by the two substations, which have on-load tap changers
(OLTCs) to keep secondary voltage at rated value. Assume that
the connection point to the transmission grid is strong. Voltage
collapse can be initiated by trip of lines or gradual increase of
load at sub-transmission level. If we assume that the secondary
voltage at the substations (at sub-transmission side) is at rated
value because of OLTC operation, the two substations function
as infinite buses, feeding the sub-transmission grid. Clearly, it
is not needed to consider the transmission side and the rest of
system behind these substations when we focus on monitored
sub-system.
B. Simplification of load area
In the study area, generally there are some radial branches
that solely draw power. If these branches are not subjects for
voltage collapse monitoring, their topology and load
impedances are not necessarily integrated into the admittance
matrix. In this case, it is sufficient to install one PMU at the
sending end to measure load impedance of the whole branch
behind. For example, at the bus #1 in Fig. 6, only one PMU is
needed to represent the load area behind. This technique
B3
B1
B2
G1
G2
study area
S1
S2
B1
B2
G1
G2
study area
Figure 5. Example of boundary selection.
Fig. 8 depicts the Thevenin impedance, load power and
impedance when the load is projected to maximum power
transfer. In this simulation, generator G2 produces 10MW and
it is able to keep its terminal voltage constant, G1 covers active
power demand from the load. The two impedances ZL1 and ZL2
are 60j and 50j () respectively; voltage level is 110kV. From
the Thevenin theorem, the Thevenin impedance is ZL1 in
parallel with ZL2. As can be seen from Fig. 8, at the instant the
load reaches its peak, load impedance is much higher than the
Thevenin impedance. Obviously, estimation of maximum
power transfer based on the Thevenin impedance is no longer a
suitable approach.
160
300
140
250
120
Load pow er
200
100
Load impedance
150
Thevenin impedance
80
100
60
50
40
G1
0
10
20
30
40
Impedance (Ohm)
B5
B4
Now, let us replace the two voltage sources V1 and V2 by
the two generators G1 and G2 respectively. If these generators
can keep the terminal voltage constant, have unlimited power
and behave like voltage sources, the estimation result is the
same as the previous case. However, if G2’s rated power is
much smaller than that of G1, the equivalent impedance
obtained from the Thevenin theorem is no longer valid to
estimate maximum power transfer.
Load power (MVA)
reduces the size of the admittance matrix of the concerned subsystem.
0
50
Time (s)
Figure 8. Load power, impedance and Thevenin equivalence without
modification.
Sub 2
400kV
Sub 1
PMU
#1
132kV
study area
Figure 6. An example of sub-transmission system and simplification of load
area.
V.
IMPACT OF ACTIVE POWER DISPATCH
A. An example of a simple system
Let us consider an electric circuit shown in Fig. 7. The load
represented by the impedance ZLoad is fed by the two voltage
sources V1 and V2. When the load grows, the currents of V1
and V2 will increase to meet the new demand. The share of this
additional power is determined by the two impedances ZL1 and
ZL2. If the load is kept increasing, the maximum power transfer
will occur when the load impedance is equal to ZL1 in parallel
with ZL2. In this case the Thevenin impedance and power
matching theory produce correct estimation of maximum
loadability.
ZL1
V1
ZL2
ZLoad
Figure 7. An electric circuit with two voltage sources.
V2
One can notice that the factor that limits the active power
generated by G2 is not the impedance ZL2 but the rated power
of the generator itself. In order to continue using the Thevenin
equivalence-based method, we propose to model the limitation
of active power of G2 as a reactance. The idea is to find a
fictitious reactance based on G2 rated power, connecting G2
and the load. This reactance replaces the physical impedance
ZL2. After that the proposed method in Section III can be used
to estimate the Thevenin impedance at the monitored bus. The
modelling approach is presented in the next paragraph.
Consider the circuit in Fig.1, where ZTh now becomes a
resistance RTh and the load is resistive, represented the
resistance RL. Assume that the maximum loadability is Pmax. At
the maximum power transfer point, RTh and RL are equal. Then
it is simple to compute RTh by
RTh 
ETh2
4 Pmax
(13)
Next, pertaining Fig. 7, we replace ZL2 by a fictitious pure
reactance XG2, which represents the rated power of the
generator. We propose to use the value of RTh in (13) to
determine the size of XG2 by
X G2 
V22
4 PG 2
(14)
where V2 and PG2 are the rated voltage and active power of
generator G2. The modified Thevenin impedance is now ZL1 in
parallel with XG2.
With this adjustment, the modified Thevenin impedance
along with load impedance and power of the previous
simulation is shown Fig. 9. It can be seen that better estimation
is obtained. The error of the estimated impedance is about 3
(6.34%). However, this error results in a minor mismatch in
maximum power estimation as depicted in Fig. 10.
160
Yeq 
YIeq
where YIeq is a 3x3 equivalent matrix, which is quite dense.
#2
#3
#4
#5
#6
#7
#1
300
140
250
120
200
Load pow er
Load impedance
150
Thevenin impedance
100
80
100
60
40
50
0
10
20
30
Impedance (Ohm)
Load power (MVA)
#8
0
50
40
Time (s)
Figure 9. Load power, impedance and modified Thevenin equivalence.
Power (MVA)
150
100
Estimated
Load
50
0
0
10
20
30
40
50
Time (s)
G1
Figure 11. A small 8-bus meshed grid.
Based on YIeq, we can construct a simplified system with
three retained buses as seen in Fig. 12. It is noticed that there
are three impedances that are of interest; they are Z18, Z88 and
Z78. Since G1 is infinite bus, Z18 functions as the constraint to
limit power transfer from G1 to bus #8. The impedance Z88 is
the equivalent load allocated to bus #8. This impedance is
important to the Thevenin impedance estimation since it
contains loading condition of the simplified sub-system. The
Z78 is the coupling impedance between bus #7 and #8; it is used
to determine the virtual power flow between bus #7 and #8,
which is further described in next paragraph. The other
impedances are the result of simplification process, and they do
not affect the estimation of the Thevenin impedance seen at bus
#8; therefore they are neglected.
Figure 10. Load power versus estimated maximum loadability after
modification.
#1
B. An example of a meshed system
The idea is to retain all generators and the concerned load
bus and simplify the rest. Assume that bus #8 in Fig. 11 is the
monitored bus; generator G1 represents an infinite system and
G2 is a small generator. As presented in Section III, we can
establish the modified 8x8 admittance matrix as
YEB 

YI 
(15)
where YI contains all generators and the considered bus (#1, #7
and #8); YE contains the rest of the system that we exclude and
YEB is the coupling sub-matrix of the two sub-systems.
Applying Gaussian elimination technique on Yeq to
eliminate YEB below the diagonal until it yields
#8
Z18
G1
In meshed power systems, power from generators is
diverted through various paths to serve different loads. To
assess voltage stability at one particular load bus, it is not
viable in real time to determine the origin of incoming power.
There is no direct link between generators and the considered
bus as the case we examined above. Therefore, it is necessary
to establish direct connection between them. This obstacle is
overcome by a system simplification technique that is coming
in the next paragraph.
Y
Yeq   E
 YEB
G2
#7
G2
Z78
Z88
Figure 12. Equivalent meshed grid after Gaussian elimination.
To evaluate voltage stability at bus #8, there are two cases
regarding the power flow between bus #7 and #8 in the new
simplified system:
-
Power flowing from bus #8 to #7 (Case 1): remove bus
#7 and impedances connected to this bus. Add new
load impedance Z87new connected to bus #8, which
extracts the same power from bus #8. The final
reduced system is shown in Fig. 13.
-
Power flowing from bus #7 to #8 (Case 2): the system
is similar to the one in Fig. 7. Therefore, (14) is used to
determine the impedance Z78new that limits the power
transferred from bus #7 to bus #8, in which V2 is the
rated voltage and PG2 is the virtual calculated power
injected to bus #8 from bus #7. The final reduced
system is shown in Fig. 14.
After above simplification process, it is possible to apply
proposed method presented in Section III to assess voltage
stability at the monitored bus. Specifically, the Thevenin
impedance in Case 1 is Z18 in parallel with Z88 and Z87new;
meanwhile it is Z18 in parallel with Z88 and Z78new in the Case 2.
#1
#8
Z18
VI.
G1
Z88
Z87new
Figure 13. Equivalent grid when power flows from bus #8 to #7.
#1
Z18
#7
Z78new
#8
G1
G2
Z88
Figure 14. Equivalent grid when power flows from bus #7 to #8.
50
40
30
25
With modification
20
Thevenin impedance
20
Load impedance
15
10
5
0
Load
0
5
Unmodified Thevenin equivalence
0
10
20
30
Time (s)
40
50
10
15
Time (s)
60
Figure 15. Load power at bus #8 and estimated maximum loadabilities based
on modified and unmodified Thevenin impedance estimation.
Figure 17. Estimated Thevenin impedance versus the monitored load
impedance, scaling only considered load.
0.08
1
5000
0
4000
-1
3000
-2
2000
-3
-4
-5
0
0
10 10
20 20
30 30
Time (s)
4040
1000
Power
Impedance
0
50
50
60
Fictitious impedance (Ohm)
Power flow from #7 to #8 (MW))
In the first test, the load at monitored bus is classically
increased until it crosses the maximum power transfer point.
All the loads are treated as constant power. The impedances
and power obtained from the load and estimation scheme are
plotted in Fig. 17 and Fig. 18. At t = 11.5s, the load hits the
maximum loadability, which is also equal to the maximum
power transfer estimated from the proposed method; therefore
the considered load is no longer on the rise, but it instead starts
going down. At this instant, the load impedance, which is in
decline, meets the estimated Thevenin impedance seen at this
bus. Obviously, the method shows quite good estimation of the
Thevenin impedance and maximum loadability.
Impedance (pu)
60
SIMULATION RESULTS
The study sub-system is part of the Norwegian transmission
system. This 30-bus area has voltage level of 132kV, supplied
by a 420/132kV substation and a large power plant connected
to the 132kV grid. These sources do not have limits in meeting
the load demand; they share quite equally load burden in this
area. However, long 132kV lines impose constraint on
maximum power transfer, and consequently voltage collapse.
In this case study, only 30 buses in this area are taken into the
admittance matrix to estimate the Thevenin impedance, but the
simulation in PSS/E is run on the model of the whole
Norwegian transmission system. The illustration in this article
is to monitor voltage stability at the weakest bus in the study
area so that only one PMU is needed to measure phasors of
load voltage and current of this concerned bus. The estimation
of the Thevenin impedance is based on load power of other
buses, which can be obtained from state estimator in practice.
Apparent power (pu)
Apparent power (MVA)
To illustrate the aforementioned proposed approach, the test
is conducted on the system shown in Fig. 11. Generator G2
produces only 5MW, and G1 covers all load demand. Like the
classical analysis, the load at bus #8 modelled as constant
impedance rises until it hits the power limit. Fig. 15 depicts the
load power and the estimated maximum power. The green
curve is the maximum loadability estimated from the method
proposed in Section III, not taking the impact of active power
dispatch into account. Since active power of G2 is much
smaller than that of G1, this algorithm fails to estimate the
maximum load power. The red curve however shows better
estimation since it utilizes the modification described above. In
this case, G2 is not considered as bulk power source. The
fictitious power flow bus #7 and bus #8 is computed in real
time and decides the size of the fictitious impedance. They are
depicted in Fig. 16.
10
there are small generators, it is needed to adjust the estimation
approach as presented in this section.
0.07
0.06
0.05
0.04
Estimated max loadability
Load
0
5
10
15
Time (s)
Figure 16. Fictitious power flow from bus #7 to #8 and impedance Z87new.
In summary, if generators in the study area contribute their
power quite equally to the demand. The modification described
in this section is not needed; the method proposed in Section III
is sufficient to estimate the Thevenin impedance. Otherwise, if
Figure 18. Estimated maximum power transfer and the monitored load power,
scaling only considered load.
In the second test case, all the loads in the study area are
expanded. They are modeled as constant power. By monitoring
the load and Thevenin impedances during the load expansion,
Impedance (pu)
25
10
5
10
15
Time (s)
Figure 19. Estimated Thevenin impedance and the impedance of monitored
load, expanding all loads in the study area.
Apparent power (pu)
0.08
0.07
0.06
0.05
0.04
Estimated max loadability
Load
0
5
10
15
Time (s)
Figure 20. Estimated maximum power transfer and the monitored load,
expanding all loads in the study area.
In the third case, one of the important lines carrying about
120MW is tripped. All the loads in the study area are modeled
by the model CLOD in PSS/E, which is comprised of 80% of
large motor, 20% of small motor and 20% of constant power
load. After losing the line at t = 2s, all the loads try to restore
their power. That leads to voltage collapse at t = 4.7s at the
weakest bus as observed in Fig. 21 and Fig. 22. Due to the
topology change at t = 2s, the Thevenin impedance seen at the
concerned bus varies slightly from 6.8 pu to 6.9 pu. During the
load restoration period, the Thevenin impedance has a minor
drop due to the decreasing impedance of the other loads in the
study area. Interestingly, the estimated maximum loadability
also falls sharply and meets load power just before the voltage
collapse. In this case, the magnitude of the Thevenin
impedance is not a decisive factor that determines the strength
(maximum loadability) of the system.
Impedance (pu)
30
Thevenin impedance
Load impedance
25
20
6.9
6.8
6.7
1.9
15
2
5
0.5
1
1.5
2
2.5
Time (s)
3
3.5
4
4.5
Figure 21. Estimated Thevenin impedance versus the monitored load
impedance, after tripping one heavily loaded line.
0.02
0
0.5
1
1.5
2
2.5
Time (s)
3
3.5
4
4.5
5
It is noted that the Thevenin equivalence is no longer a
suitable method when generators behave differently from
voltage source, especially in terms of active power limit. A
new impedance whose function is similar to that of Thevenin
impedance in introduced to calculate maximum power transfer.
Under this circumstance, PMU installations at small generators
are needed in addition to concerned load buses.
Finally, the proposed method has been validated by
simulations on the simple 8-bus system and the Norwegian
transmission system.
REFERENCES
[1]
[2]
[3]
[4]
[6]
Variation of Thevenin impedance at t = 2s
0
0.04
Figure 22. Estimated maximum power transfer and the monitored load power,
after tripping one heavily loaded line.
[5]
2.1
10
0.06
This paper has presented a methodology to assess voltage
stability at load buses based on estimation of the Thevenin
impedance and the maximum loadability. Based on the
information of system topology and load power, which are
available from existing state estimator, the Thevenin
impedance is estimated. Therefore, PMU installation is needed
only at concerned buses to compute the maximum loadability,
which functions as the main indicator for voltage stability
monitoring. It has been demonstrated that it is sufficient to take
into calculation only part of the system that is vulnerable to
voltage instability. This makes the proposed method feasible
for practical applications, especially for sub-transmission
system where load are normally supplied by strong connection
to the transmission level and voltage stability problem is
imposed by the power lines.
Load impedance
5
Load
0.08
VII. CONCLUSION
15
0
Estimated max loadability
0.1
0
Thevenin impedance
20
0
Apparent power (pu)
voltage collapse has been detected at the weakest bus;
meanwhile the rest is still within its limit. Fig. 19 and Fig. 20
present the performance of the proposed algorithm at the
concerned bus. It is noticed that during the load increase
process the Thevenin impedance drops slightly since load
impedances of the whole system are decreasing; however the
estimated maximum loadability shows a pretty large drop. This
is the impact of modelling the load as constant power. In this
case, the actual maximum power transfer is revealed when the
load is approaching its limit.
5
[7]
V. Ajjarapu and C. Christy, “The continuation power flow: a tool for
steady state voltage stability analysis,” IEEE Trans. Power Syst., vol. 7,
pp. 416-423, Feb. 1992.
B. Gao, G. K. Morison, P. Kundur, “Voltage stability evaluation using
modal analysis,” IEEE Trans. Power Syst., vol. 7, pp. 1529-1542, Nov.
1992.
K. Vu, M. M. Begovic, D. Novosel and M. M. Saha, “Use of local
measurements to estimate voltage-stability margin,” IEEE Trans. Power
Syst., vol. 14, pp. 1029-1035, Aug. 1999.
S. Corsi and G. N. Taranto, "A real-time voltage instability identification
algorithm based on local phasor measurements," IEEE Trans. Power
Syst., vol. 23, pp. 1271-1279, Aug. 2008.
Y. Wang, I. R. Pordanjani, W. Li, W. Xu, T. Chen, E. Vaahedi and J.
Gurney, "Voltage stability monitoring based on the concept of coupled
single-port circuit," IEEE Trans. Power Syst., vol. 26, pp. 2154-2163,
Nov. 2011.
J. J. Grainger and W. D. Stevenson, Power System Analysis, Singapore:
McGraw-Hill, 1994, pp. 283-294.
M. Begovic, B. Milosevic and D. Novosel, “A novel method for voltage
instability protection,” in Proc. 2002 IEEE Computer Society System
Sciences Conf., pp. 802-811.