You are on page 1of 10

IEEE TRANSACTIONS ON SMART GRID, VOL. 3, NO.

3, SEPTEMBER 2012

1427

A New Control Strategy to Mitigate the Impact of


Inverter-Based DGs on Protection System
Hesam Yazdanpanahi, Student Member, IEEE, Yun Wei Li, Senior Member, IEEE, and Wilsun Xu, Fellow, IEEE

AbstractDespite their undoubted advantages, Distributed


Generation (DG) systems can negatively impact some aspects of
the distribution system operation. In this paper, impacts of inverter-based DGs on fuse-recloser coordination in the fuse-saving
protection scheme are thoroughly studied. Various fault conditions with different fault resistances and the effects of different
DG locations are investigated. Also, the effects of DG reactive
power injection, known as a DG potential ancillary service, on
the protection scheme are studied. Furthermore, in order to
mitigate the impact of DG on the protection coordination, a simple
and effective control strategy is proposed. This strategy limits
the DG output current according to the DG terminal voltage.
Extensive simulations at different fault conditions and different
DG penetration levels showed that the proposed control method is
able to eliminate DGs contribution during the fault, and consequently eliminate its impact on the fuse-recloser coordination. In
comparison to other methods, this strategy is inexpensive, easy to
implement, does not limit DG capacity during normal condition,
and does not require any change in the original protection system.
The simulation results also demonstrated that the proposed
method is robust against non-fault transient disturbances such as
load switching, starting of induction motors, etc.
Index TermsDistributed generation (DG), distribution system
protection, fuse-recloser coordination, power electronics interface.

I. INTRODUCTION

ISTRIBUTED generation (DG) provides several advantages for utilities. First, DG is an alternative for satisfying
the increasing load demand without the need for transmission
system expansion [1]. Since DG is directly connected to the distribution network, the transmission loss through the traditional
central generation can be reduced. Second, DG enables renewable energy sources such as PV, wind, and fuel cells to be integrated into the power network [2]. Such integration addresses
the increasing concerns on the environment, energy costs, and
energy security. Finally, DGs can provide ancillary services to
utilities, including voltage regulation, reactive power compensation, and active power filtering [3][5].
However, DG implementation can also negatively impact
utilities, especially their distribution protection system. For example, the DGs unintentional islanding operation may damage
network elements and is also a safety concern for utilities
Manuscript received September 28, 2011; revised December 21, 2011; accepted January 10, 2012. Date of publication July 20, 2012; date of current version August 20, 2012. Paper no. TSG-00561-2011.
The authors are with the Department of Electrical and Computer Engineering, University of Alberta, 2nd floor, ECERF, 9107-116str., Edmonton,
AB T6G 2V4, Canada (e-mail: yazdanpa@ualberta.ca; yunwei.li@ualberta.ca;
wxu@ualberta.ca).
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org.
Digital Object Identifier 10.1109/TSG.2012.2184309

[6][10]. Sympathetic tripping of overcurrent protection relays


has been considered as another negative effect of DG. In this
case, when a fault happens on a feeder adjacent to the DGs
feeder, the DG may feed the fault current. Since traditional
protection systems have been designed based on the assumption of a unidirectional current, the DG feeders protection will
operate, disconnect the DG feeder from the substation, and
consequently interrupt power delivery to the DGs feeder even
though it has no fault on it [11][14].
DG can also negatively impact the fuse-recloser miscoordination in the fuse-saving protection scheme. In this scheme, the
feeders recloser is designed to operate sooner than the fuse
to prevent fuse damage during temporary faults. However, the
DGs contribution in a fault condition may increase the fuse current and make it operate faster than the recloser, leading to the
failure of the fuse-recloser coordination [15][18].
To mitigate DGs impacts on the traditional protection
scheme, several methods have been proposed, which can be
classified into four major categories:
1. Limiting DG capacity: these methods try to find the DGs
maximum capacity, so that the DG does not impact the
protection system [19][23].
2. Modifying the protection system: these methods are based
on using extra breakers or reclosers, reconfiguration of
the network, or using distance or directional relays, which
are not commonly used in distribution system protection
[24][27].
3. Using adaptive protection: in contrast to the traditional protection, in adaptive protection, relays are capable of communicating with each other and have access to remote measurements. In addition, their settings can be dynamically
adjusted by using a fast processing unit [28][31].
4. Utilizing fault current limiters (FCLs): FCL devices are series elements that show negligible impedance during the
networks normal operation. In contrast, during the fault
condition, their impedances increase immediately to restrict the current flow through their branches [32][35].
While effective for mitigating the DG impacts on the protection system, these solutions have some obvious disadvantages.
For example, limiting the DG capacity is not a desirable solution since doing so also limits the DG penetration level. Modifying the protection system is costly. In addition, this method
requires utility involvement and makes the protection procedure
more complicated. Similarly, adaptive protection requires new
systems like communication infrastructures and fast processing
units. Finally, utilizing FCLs is also undesirable due to the additional cost for utility or DG owners.
The main objective of this paper is to propose a simple control strategy for inverter-based DGs to solve protection-related

1949-3053/$31.00 2012 IEEE

1428

IEEE TRANSACTIONS ON SMART GRID, VOL. 3, NO. 3, SEPTEMBER 2012

Fig. 1. Fuse-recloser protection scheme in distribution feeders.

problems without restricting DG utilization during normal conditions. The focus of the paper is on the fuse-recloser coordination, which is one of the most important protection schemes in
distribution systems. The impact of inverter-based DG on protection coordination has been neglected in a few studies with
the consideration of limited fault current capacity and quick inverter control response to a fault [23], [36], [37]. However, inverter-based DG systems can cause miscoordination in cases of
high DG penetration or tight protection coordination [38]. The
aggregate contributions of many small DG units or a few large
units are sufficient to increase shortcircuit level, at least for the
first few cycles, which may lead to fuse-breaker miscoordination [37], [39][43].
In this paper, the impacts of inverter-based DG systems on
fuse-recloser coordination are first thoroughly studied. Various
fault conditions and the effects of different DG locations are investigated. The effects of DG reactive power injection on the
protection scheme are also analyzed. To limit the DG impacts,
a control strategy that limits the DG output current according to
the DG terminal voltage is proposed. In contrast to the previous
methods, the proposed method is relatively inexpensive and is
applied at the DG side. As well, this methods implementation
requires no utility involvement. More importantly, no change is
required in the original protection system. In this paper, extensive simulations with different fault conditions and DG penetration levels were conducted. System disturbances such as those
caused by starting of a large induction motor were also tested.
The results showed that the proposed strategy is robust against
non-fault disturbances in the network.
II. IMPACT OF DG ON FUSE-RECLOSER COORDINATION
In the fuse-saving scheme, when a fault occurs on a lateral
like the one shown in Fig. 1, the recloser first operates once
or more times based on its fast time-current curve. Most of
the faults in the distribution system are temporary and will be
cleared during fast reclosing actions [21]. If the fault is quasipermanent, the fuse is supposed to clear the fault instead. The
time-delayed operation of the recloser will occur if the fuse fails
to interrupt the fault current. Consequently, in case of a permanent fault, the fuse is set to melt between fast and time-delayed
operation of an automatic recloser. Applying the fuse-saving
scheme has two main advantages [44]:
1. No interruption in power delivery occurs due to temporary
faults.
2. Fuse burning and replacement are needed only if the fault
is quasi-permanent.
For proper coordination, the fuse and recloser curves are selected and set in a way that for all possible faults, fuse and recloser fault currents remain within the limit shown in Fig. 2.

Fig. 2. Time-current characteristic curves of recloser superimposed on fuse


curve.

Fig. 3. Impact of DG on operation of protection system under low impedance


fault situation.

Fig. 4. Equivalent circuit of a distribution system with inverter-based DG,


during a downstream fault.

Nonetheless, insertion of DG changes the fault current experienced by the fuse and recloser. For example, for low-impedance
faults, adding DG to the system may increase the fault current
experienced by the fuse. This will push the fuse current to the
right side of point as shown in Fig. 3. In this case, the fuse
will melt simultaneously or faster than the operation of recloser,
and an undesirable permanent interruption will occur on the lateral, even for temporary faults.
For further analysis, Fig. 4 shows an equivalent circuit
for a system with an inverter-based DG when a fault occurs
downstream of the DG. In this figure, the network upstream
of the distribution substation is modeled as an equivalent ideal
voltage source , and an equivalent impedance
. In this
model, loads are neglected (considered as open circuits) during
the fault. In addition,
and
are feeder impedances from the
substation to the PCC, and from the PCC to the fault location,
respectively.
represents the recloser, and
is the fault
resistance.
In this system, the three-phase short circuit per-unit current
before implementing DG can be estimated as
(1)

YAZDANPANAHI et al.: A NEW CONTROL STRATEGY TO MITIGATE THE IMPACT OF INVERTER-BASED DGS

1429

After adding the DG, the currents through recloser and fuse
can be obtained as

(2)

(3)
At point of Fig. 2, where a low-impedance fault occurs near
the PCC
, the recloser and fuse currents
can be approximated as
(4)
(5)
From (4) and (5), it can be concluded that adding DG increases the fuse current in low-impedance faults, while the fault
current experienced by the recloser is almost constant. Thus, in
this case, the fuse may operate faster than the recloser, leading to
a failure of coordination. Suppose that before adding DG, Point
of Fig. 2 is reached for a fault with resistance
. After
adding DG, point will be reached for a fault with resistance
, where necessarily
. Consequently, fuse-recloser protection cannot be applied for faults with resistances
between
and
.
On the other hand, for high-impedance faults
the recloser and fuse currents can be estimated as

Fig. 5. Voltage and current vector diagram of: (a) DG provides active power;
(b) DG provides reactive power.

III. IMPACT OF REACTIVE POWER INJECTION


FUSE-RECLOSER COORDINATION

ON

The provision of reactive power has been considered as one of


the most important ancillary services offered by inverter-based
DGs [4], [5], [45]. However, if not controlled and coordinated
properly, this reactive power injection may worsen the protection miscoordination. In this section, the effect of DG reactive
power injection on the fuse-recloser coordination is studied.
Consider a low-impedance downstream fault as an example,
where the fuse current has two parts as shown in (5). In active
power injection from the DG, the first part of the fuse current
is about 90 out of phase with the
PCC voltage, while the second part (the DG current)
is
forced to be in phase with the PCC voltage. As a result, the two
terms have a phase shift of around 90 , and the relationship in
(8) can be obtained:

(6)

(8)

(7)

However, in reactive power injection from DG, both terms in


the fuse current (
and
) are about
90 out of phase with the PCC voltage. Consequently, the relationship in (9) is obtained:

Equations (6) and (7) show that for high-impedance faults


(like point of Fig. 2), the current experienced by the recloser
is reduced after adding DG, while the fuse current is almost constant. Under this condition, the fuse operation time remains constant, but the recloser time-delayed operation occurs with an additional delay. However, this delay will not cause miscoordination since the fuse operates sooner than the reclosers time-delayed operation.
Furthermore, analysis on the situation of a fault occurs upstream to the DG has also been conducted in a similar way. It
has been found that the situation matches closely with that of
a downstream fault. This means the location of faults does not
play a significant role in the fuse-recloser coordination. Instead,
the DG current and fault impedance will have more impacts
on the coordination. Moreover, compared to the situation of a
high-impedance fault, where the coordination is usually maintained, a low-impedance fault tends to cause more problems.
For this reason, low-impedance faults are considered in the rest
of this work.

(9)
For further illustration, Fig. 5 shows the vector diagram of
the above voltages and currents (note that the direction of DG
current is to the PCC). Fig. 5(a) and 5(b) show that the fault
current through the fuse is larger when the DG provides reactive
power. This means that in comparison to active power injection,
reactive power injection worsens the fuse-recloser coordination
situation.
In some preliminary grid codes such as IEEE 1547 [48], DG
systems are supposed to be disconnected from the grid quickly
during disturbances, so that they have no effect on system operation. However, by increasing DG penetration in networks, grid
codes are being modified [49], and in some high and extra high

1430

IEEE TRANSACTIONS ON SMART GRID, VOL. 3, NO. 3, SEPTEMBER 2012

TABLE I
IEEE STD. 1547 REQUIRED RESPONSE TO ABNORMAL VOLTAGE CONDITIONS

voltage grid connection standards like [50], [51], distributed resources are forced to stay connected to the grid during the fault,
and support the grid voltage by injecting reactive power. Nowadays, such re-evaluation is going to be extended to mediumvoltage
grids as well [52], [53],
where the DG is required to stay connected and provide reactive
power for network during voltage sag. Owing to the fact that the
fuse saving scheme is a commonly used protection method in
these medium-voltage networks, satisfying these new requirements by DG could worsen the miscoordination problem. So,
compatibility of new codes for DG interconnection to grid with
over-current protection system should be carefully taken into
account. Also, the development of DG control strategy that can
satisfy both requirements would be very desirable.
IV. THE PROPOSED INVERTER CURRENT CONTROL STRATEGY
To mitigate the impacts of DG on the fuse-recloser coordination, a simple and effective DG current control strategy is proposed in this section.
In common practice, the inverters current during a
low-impedance fault is restricted up to twice its nominal
current [54]. The ideal way to eliminate the impact of DG
on the protection system is to detect the fault and trip all the
converters in that protection zone. However, converters are
unable to differentiate between fault conditions and short-term
disturbances such as load switching. Therefore, converter
tripping in all abnormal conditions may lead to unnecessary
power-delivery interruptions.
One way to ride through short-term disturbances and avoid
excessive nuisance tripping is to introduce allowed time delay,
as is shown in Table I. If the abnormal condition remains after
the delay, according to IEEE std. 1547 [48], the converter should
be tripped and then reconnected 5 min after the system has returned to its normal condition.
Nowadays, fast automatic reclosers open and reclose in less
than 6 cycles after fault occurrence, so the above time delays
may not be effective; i.e., during the reclosing, DG still contributes to the arching fault. In addition, decreasing these delays may cause excessive nuisance tripping due to short-term
disturbances.
One idea to simultaneously solve the miscoordination
problem and ride through short-term disturbances is to reduce
the converter current according to the severity of the abnormality instead of completely blocking the converter. In this
case, the converters near the fault location, which will produce
the greatest impact on the protection system, experience the
most voltage deviation from the normal boundaries and should
significantly decrease their fault current contribution. On the

Fig. 6. Proposed strategy to determine inverter reference current.

other hand, the more distant DG units, which have no substantial effect on the protection system, can continue their power
delivery.
To implement the above-mentioned current-control strategy
according to the DG terminal voltage, the DGs reference current can be determined as in (10)
for

(10)

for
is the converter reference current,
is the maxwhere
imum output current that happens at
(the
lower boundary for normal operation according to Table I),
is the rms voltage at the DG connection node,
is the output desired power and and are constants to be
determined.
Generally the value of determines the sensitivity of the control scheme to a voltage change. A larger value of leads to
more obvious output current reduction with a voltage sag. However, too large will cause the control scheme over sensitive to
even a small voltage disturbance. With the above consideration,
is selected in this work. Once the value of is chosen,
the coefficient can be determined in such a way that the reference current in (10) remains to be a continuous function around
; i.e., can be obtained from (11), which
gives
when
.
(11)
Fig. 6 shows the flow chart illustrating the reference current
determination procedure.
V. SIMULATION INVESTIGATION RESULTS
A. Performance During Low-Impedance Fault Condition
To investigate the ability of the proposed strategy to maintain fuse-saving coordination, a 13-node test feeder system (see
Fig. 7) [55] was constructed by using MATLAB/SIMULINK.
A recloser was mounted at the substation, and an inverter-based
DG was connected at node 645. The simple current-controlled
voltage source inverter model in [23] is considered in this work.
The DGs effect on the coordination between the recloser and

YAZDANPANAHI et al.: A NEW CONTROL STRATEGY TO MITIGATE THE IMPACT OF INVERTER-BASED DGS

Fig. 7. IEEE 13-node test feeder system [55].

1431

Fig. 10. Difference between fuse and recloser operation time after adding DG
for 0.1 fault.

Fig. 8. Fuse-recloser coordination in the simulated system.

Fig. 11. Intentional phase shift in DG current during voltage sag to support
voltage.

Fig. 9. Difference between fuse and recloser operation time after adding DG
for 0.01 fault.

Fig. 12. Difference between fuse and recloser operation time for 0.01
with pi/2 phase shift in inverter current.

the fuse on 645646 was studied for faults that occur in the
middle of 645646. Fig. 8 shows the time-current characteristic
curves of the recloser and fuse used in simulations.
In the simulated system, point occurs for fault resistance
of 0.01 , and point occurs for fault resistance of 11.5 .
In other words, the fuse and recloser operate properly for fault
resistances between 0.01 to 11.5 . For faults with resistance
lower than 0.01 , the fuse will operate faster than the recloser,
and the use of fuse-saving scheme is not feasible.
Fig. 9 shows the consequence of adding DG at different penetration levels
on the
overcurrent coordination when a 0.01 fault occurs. As Fig. 9
reveals, after adding DG even at low penetration levels, the fuse
operates faster than the recloser, and the protection coordination
is lost.
This problem will appear in situations with even higher fault
resistances. Fig. 10 shows that miscoordination occurs in a fault
with 0.1 resistance in the presence of DG at high penetration
levels. As shown, although the fault resistance is higher than
the one in point , adding DG at penetration levels higher than
70% will cause miscoordination.

To investigate the impact of reactive power injection on the


protection coordination and confirm the analysis of the previous
sections, the DG control scheme was modified so that the DG
provided active power during normal conditions, but injected a
fully reactive current when it experienced a voltage sag
. Fig. 11 shows the PCCs voltage magnitude and the
DG current phase angle for a fault occuring at
. Fig. 12
shows the difference between fuse and recloser fast operation
times under this control scheme, for a 0.01 fault.
The comparison between Figs. 9 and 12 reveals that the fuse
melted sooner when the DG injected reactive power. This impact occurred even in situations with higher fault resistances.
Fig. 13 shows that with reactive power injection, miscoordination occured for
even at low penetration levels like
30% (compared to 70% in Fig. 10). In addition, Fig. 14 shows
that reactive power injection caused miscoordination in 0.2
faults for penetration levels higher than 75%. Note that no miscoordination occurred for 0.2 faults at any penetration level
when DG provided only active power.
Finally, Fig. 15 shows the effect of the proposed control
scheme on the protection coordination. As this figure shows,

fault

1432

IEEE TRANSACTIONS ON SMART GRID, VOL. 3, NO. 3, SEPTEMBER 2012

Fig. 13. Difference between fuse and recloser operation time for 0.1
with pi/2 phase shift in inverter current.

fault

Fig. 14. Difference between fuse and recloser operation time for 0.2
with pi/2 phase shift in inverter current.

fault

Fig. 16. PCC voltage for 0.01

fault with DG at 30% penetration level.

Fig. 17. DG output current for 0.01

Fig. 18. Fuse current for 0.01


Fig. 15. Difference between fuse and recloser operation time for 0.01
with inverter current reduction control.

fault with DG at 30% penetration level.

fault with DG at 30% penetration level.

fault

this control scheme successfully solved the miscoordination


in low-resistance faults due to DG injection. This figure also
reveals that with the proposed control scheme, the DG penetration level had almost no effect in low impedance faults.
This is due to the fact that PCC voltage is very low under a
low-impedance fault (see Fig. 16), and as shown in Fig. 17, the
DG output current decreased to almost zero with the proposed
method. As a result, the protection devices did not experience
any difference between this system with DG and the previous
system with no DG, and the coordination was maintained (see
Fig. 18). Note that this result contrasts with the case when traditional control was applied to the inverter. In this case, during
voltage sags, traditional control increased the inverter current
until it reached the preset current threshold
, which is
normally set to
the
current. Therefore, the
fuse current was increased, and consequently, miscoordination
occurred.
In the next step, in order to show the effectiveness of the proposed strategy in multiple DG cases, two DG units were used in
the test system. One was implemented at Node 634 with a capacity equal to 20% of total load of the system. The other was

Fig. 19. Output current of inverter installed at Node 634.

installed at Node 675 with a capacity equal to 60% of the total


load of the distribution system. Their effect on the coordination between the recloser and the fuse on 645646 was studied
for a 0.1 fault occurring in the middle of 645646. Figs. 19
and 20 show these inverters output currents when controlled
by the proposed strategy. In addition, Fig. 21 demonstrates the
fuse current in both the traditional and current-restricting strategies. As this figure reveals, in contrast to the traditional control,
the proposed strategy successfully kept the fuse current below
the margin, and consequently maintained the coordination even
when multiple DGs with high penetration at different locations
were used.

YAZDANPANAHI et al.: A NEW CONTROL STRATEGY TO MITIGATE THE IMPACT OF INVERTER-BASED DGS

Fig. 20. Output current of inverter installed at Node 675.

Fig. 21. Fuse current for 1

1433

Fig. 23. Inverter output power for 8-

fault in the middle of 645646.

fault with DG units at Nodes 634 and 675.

Fig. 22. Output current of inverters for 2-

Fig. 24. PCC voltage for 8-

fault in the middle of 645646.

Fig. 25. Fuse current for 8-

fault in the middle of 645646.

fault in the middle of 645646.

B. Performance During High-Impedance Fault Condition


The converters operation controlled by the proposed strategy
during medium- and high-impedance faults is studied in this
stage of the simulations. In the first case, three DGs at a 30%
total penetration level (10% each) were implemented on Nodes
633, 645, and 675, and a 2- fault was simulated in the middle
of 645646. Fig. 22 shows the converters output current. This
figure reveals that the more distant converters experienced
lower voltage sags, and consequently, provided more current
during a fault.
In the second case, a DG has been installed at Node 634 with
a capacity equal to the spot load at this node, and an 8- fault
has been simulated from 0.2 s to 0.3 s in the middle of Nodes
645646. Fig. 23 shows the inverter output power in this situation. Fig. 24 shows the PCC voltage and Fig. 25 shows the
fuse current. According to these results, the converter with the
proposed control strategy can successfully ride through high
impedance or distant faults without losing its power delivery
capability, and improve the voltages of local nodes without imposing any impact on the protection system.

Fig. 26. A distribution network with fuse-saving protection.

C. Performance During Other Disturbances


To test the performance of the proposed control method under
a network disturbance other than a fault condition, a distribution
feeder with an induction motor load was simulated. The single
line diagram of this feeder is shown in Fig. 26. The distribution system parameters are listed in Table II.The substation was
modeled by a three-phase balanced voltage source and equivalent impedance, the line was modeled by series R-L branches.
As shown in Fig. 26, this feeder had five laterals, and a recloser
was mounted at its middle. is an 80 A fuse, and is a recloser
in conjunction with an IEEE extremely inverse relay, which is
the most suitable option for fuse-saving practice [56].

1434

IEEE TRANSACTIONS ON SMART GRID, VOL. 3, NO. 3, SEPTEMBER 2012

TABLE II
TEST SYSTEM PARAMETERS

Fig. 29. Voltage at terminals of IM during motor starting for both traditional
and proposed control strategies.

Fig. 27.

s characteristic curve superimposed on

s.

Fig. 30. Inverter output current during the motor starting transient.
TABLE III
INDUCTION MOTOR PARAMETERS

Fig. 28. DG output power during motor starting for both traditional and proposed control strategies.

Fig. 27 shows the fuses time-current curve superimposed


on the reclosers curve. Point is the fuse operation point for
a three-phase fault at the beginning of
before adding an
inverter-based DG. Since
is to the left of , the recloser
operates faster than the fuse, and
will be saved in case of
temporary faults at
. In contrast, after adding a 1-MW inverter-based DG, the fuses operation point moves from to
, and fuse saving is no longer practical unless the proposed
control strategy is utilized.
To study the response of inverter-based DG controlled by the
proposed strategy during non-fault disturbances, the starting of a
1-MW induction motor (IM) was considered and simulated. The
induction motor parameters are listed in Table III. As demonstrated in Fig. 28, only a 5% reduction occurs in the DG output
power. Such a reduction barely worsens the voltage at the PCC.
As Fig. 29 reveals, the time-voltage curves at the motor terminal match for both the traditional and current restricting control strategies. Finally, Fig. 30 shows the inverter output current
during this transient. As can be seen, the current limiting control is not enabled as the inverter terminal voltage is higher than
0.88 p.u.

These simulation results not only prove that the proposed


control strategy is appropriate for fuse saving, but also demonstrate the robustness of this strategy against non-fault disturbances. In other words, this strategy reduces inverter output only
during low-impedance faults when fuse-burning is possible. In
contrast, the DG output power remains almost constant for other
network disturbances and consequently, it does not cause further
voltage reduction.
VI. CONCLUSION
Both analytical and simulation studies showed that inverterbased DGs even at low-penetration levels can cause miscoordination between the fuse and the recloser in fuse-saving practices. For low-impedance faults, DG increased the fault current
through the fuse, made it burn faster than expected, and consequently affected the fuse-recloser coordination. In addition,
it was demonstrated that reactive power injection by DG worsened the situation, and miscoordination occurred in higher fault
resistances and lower DG penetration levels.
To solve this problem, a simple and effective control strategy
was proposed which adjusts the inverter output current according to the severity of the abnormality; i.e., the deeper the
voltage sag experienced by the DG, the less current it produces.

YAZDANPANAHI et al.: A NEW CONTROL STRATEGY TO MITIGATE THE IMPACT OF INVERTER-BASED DGS

The simulation results showed that this scheme successfully


mitigated the miscoordination problem during the fault condition. Another advantage of this control strategy is that it is
robust enough against non-fault disturbances; i.e., it avoids nuisance reduction in DG power delivery. As an example, it was
shown that the inverter output power remained almost constant
during voltage disturbances like those caused by starting large
induction motors.
REFERENCES
[1] A. Piccolo and P. Siano, Evaluating the impact of network investment deferral on distributed generation expansion, IEEE Trans. Power
Syst., vol. 24, no. 3, pp. 15591567, Aug. 2009.
[2] J. M. Guerrero, F. Blaabjerg, T. Zhelev, K. Hemmes, E. Manmasson, S.
Jemei, M. P. Comech, R. Granadino, and J. I. Frau, Distributed generation: Toward a new energy paradigm, IEEE Ind. Electron. Mag.,
vol. 4, no. 1, pp. 5264, Mar. 2010.
[3] M. Triggianese, F. Liccardo, and P. Marino, Ancillary services performed by distributed generation in grid integration, in Proc. Int. Conf.
Clean Electr. Power, 2007, pp. 164170.
[4] J. He, Y. W. Li, and S. Munir, A flexible harmonic control approach
through voltage controlled DG-grid interfacing converters, IEEE
Trans. Ind. Electron., vol. 59, no. 1, pp. 444455, Jan. 2012.
[5] J. He, S. Munir, and Y. W. Li, Opportunities for power quality
improvement through DG-grid interfacing converters, in Int. Power
Electron. Conf. (IPEC10), in Proc. IEEE ECCE-Asia, 2010, pp.
16571664.
[6] X. Wang, W. Freitas, and W. Xu, Dynamic non-detection zones of
positive feedback anti-islanding methods for inverter-based distributed
generators, IEEE Trans. Power Del., vol. 26, no. 2, pp. 11451155,
Apr. 2011.
[7] A. H. Kasem Alaboudy and H. H. Zeineldin, Islanding detection for
inverter-based DG coupled with frequency-dependent static load,
IEEE Trans. Power Del., vol. 26, no. 2, pp. 10531063, Apr. 2011.
[8] H. Vahedi, R. Noroozian, A. Jalilvand, and G. B. Gharehpetion , A
new method for islanding detection of inverter-based distributed generation using DC-link voltage contro, IEEE Trans. Power Del., vol.
26, no. 2, pp. 11761186, Apr. 2011.
[9] H. H. Zeineldin, A Q-f droop curve for facilitating islanding detection
of inverter-based distributed generation, IEEE Trans. Power Electron., vol. 24, no. 3, pp. 665673, Mar. 2009.
[10] S. H. Lee and J. W. Park, New islanding detection method for
inverter-based distributed generation considering its switching frequency, IEEE Trans. Ind. Appl., vol. 46, no. 5, pp. 20892098,
Sep./Oct. 2010.
[11] W. El-Khattam and T. S. Sidhu, Restoration of directional overcurrent relay coordination in distributed generation systems utilizing fault
current limiter, IEEE Trans. Power Del., vol. 23, no. 2, pp. 576585,
Apr. 2008.
[12] J. C. Gmez and M. M. Morcos, Coordination of voltage sag and overcurrent protection in DG system, IEEE Trans. Power Del., vol. 20, no.
1, pp. 214218, Jan. 2005.
[13] W. El-Khattam and T. S. Sidhu, Resolving the impact of distributed
renewable generation on directional overcurrent relay coordination: A
case study, IET Renew. Power Gener., vol. 3, no. 4, pp. 415425,
2009.
[14] C. J. Mozina, A tutorial on the impact of distributed generation (DG)
on distribution systems, in Proc. Protective Relay Engineers Annu.
Conf., 2008, pp. 591609.
[15] R. A. Walling, R. Saint, R. C. Dugan, J. Burke, and L. A. Kojovic,
Summary of distributed resources impact on power delivery system,
IEEE Trans. Power Del., vol. 23, no. 3, pp. 16361644, Jul. 2008.
[16] H. Cheung, A. Hamlyn, L. Wang, C. Yang, and R. Cheung, Investigations of impacts of distributed generation on feeder protection, in
Proc. IEEE Power Energy Soc. Gen. Meet., 2009, pp. 17.
[17] R. C. Dugan, T. S. Key, and G. J. Ball, Distributed resources standards, IEEE Ind. Appl. Mag., vol. 12, pp. 2734, Feb. 2006.
[18] M. H. Kim, S. H. Lim, J. F. Moon, and J. C. Ki, Method of recloserfuse coordination in a power distribution system with superconducting
fault current limite, IEEE Trans. Appl. Supercond., vol. 20, no. 3, pp.
11641167, Jun. 2010.
[19] T. Seegers et al., Impact of distributed resources on distribution relay
protection, Rep. to Line Protection Subcommitee, Power System
Relay Committee, Power Engineering Society, IEEE, 2004.

1435

[20] S. Chaitusaney and A. Yokoyama, An appropriate distributed generation sizing considering recloser-fuse coordination, in Proc. IEEE
Transm. Distrib. Conf. Exhib.: Asia Pacific, 2005, pp. 16.
[21] S. Chaitusaney and A. Yokoyama, Prevention of reliability degradation from recloser-fuse miscoordination due to distributed generatio, IEEE Trans. Power Del., vol. 23, no. 4, pp. 25452554, Oct.
2008.
[22] J. Chen, R. Fan, X. Duan, and J. Cao, Penetration level optimization for DG considering reliable action of relay protection device constrains, in Proc. Int. Conf. Sustainable Power Gener. Supply, 2009,
pp. 15.
[23] T. K. Abdel-Galil, A. E. B. Abu-Elnien, E. F. El-saadany, A. Girgis, Y.
A.-R. I. Mohamed, M. M. A. Salma, and H. H. M. Zeineldin, Protection coordination planning with distributed generation, Canmet Energy Technology Centre, Varennes, QC, Canada, 2007.
[24] H. B. Funmilayo and K. L. Buyler-Purry, An approach to mitigate the
impact of distributed generation on the overcurrent protection scheme
for radial feeders, in Proc. IEEE Power Syst. Conf. Expo., 2009, pp.
111.
[25] F. A. Viawan, D. Karlsson, A. Sannino, and J. Daalde, Protection
scheme for meshed distribution systems with high penetration of distributed generation, in Proc. Power Syst. Conf. Adv. Metering, Protection Control, Commun., Distrib. Resources, 2006, pp. 99104.
[26] I. M. Chilvers, N. Jenkins, and P. A. Crossley, The use of 11 kV distance protection to increase generation connected to the distribution
network, in Proc. Int. Conf. Develop. Power Syst. Protect., 2004, vol.
2, pp. 551554.
[27] D. Uthitsunthorn and T. Kulworawanichpong, Distance protection of
a renewable energy plant in electric power distribution systems, in
Proc. Int. Conf. Power Syst. Technol., 2010, pp. 14.
[28] H. Wan, K. K. Li, and K. P. Wong, An adaptive multiagent approach to
protection relay coordination with distributed generators in industrial
power distribution system, IEEE Trans. Ind. Appl., vol. 46, no. 5, pp.
21182124, Sep./Oct. 2010.
[29] S. A. M. Javadian, R. Tamizkar, and M. R. Haghifam, A protection
and reconfiguration scheme for distribution networks with DG, in
Proc. PowerTech, 2009, pp. 18.
[30] J. Ma, J. Li, and Z. Wang, An adaptive distance protection scheme for
distribution system with distributed generation, in Proc. 5th Int. Conf.
Critical Infrastruct., 2010, pp. 14.
[31] M. Baren and I. El-Markabi, Adaptive overcurrent protection for
distribution feeders with distributed generators, in Proc. Power Syst.
Conf. Expo., 2004, vol. 2, pp. 715719.
[32] W. El-Khattam and T. S. Sidhu, Restoration of directional overcurrent relay coordination in distributed generation systems utilizing fault
current limiter, IEEE Trans. Power Del., vol. 23, no. 2, pp. 576585,
Apr. 2008.
[33] M. M. R. Ahmed, G. Putrus, L. Ran, and R. Penlington, Development
of a prototype solid-state fault-current limiting and interupting device
for low-voltage distribution networks, IEEE Trans. Power Del., vol.
21, no. 4, pp. 19972005, Oct. 2006.
[34] H. Yamaguchi and T. Kataoka, Current limiting characteristics of
transformer type superconducting fault current limiter with shunt
impedance and inductive load, IEEE Trans. Power Del., vol. 23, no.
4, pp. 25452554, Oct. 2008.
[35] Y. Zhang and R. A. Dougal, Novel dual-FCL connection for adding
distributed generation to a power distribution utility, IEEE Trans.
Appl. Supercond., vol. 21, no. 3, pp. 21792183, Jun. 2011.
[36] R. Dugan, R. Zavadil, and D. Van Holde, Interconnection guidelines
for distributed generation, E SOURCE & Electrotek Concepts, Aug.
2002 [Online]. Available: http://grouper.ieee.org/groups/scc21/pdfs/
ESourceDG_InterconnectionAppGuide.pdf
[37] J. Keller and B. Kroposki, Understanding fault characteristics of inverter-based distributed energy sources, National Renewable Energy
Laboratory, Golden, CO, NREL/TP-550-46698, 2010.
[38] G. Hernandez-Gonzalez and C. Abbey, Effect of Adding Distributed
Generation to Distribution Networks Case Study 3: Protection Coordination Consideration with Inverter and Machine Based DG
Canmet Energy, Varennes, QC, Canada, Report-2009-043 (RP-TEC)
411-MODSIM, 2009.
[39] B. Kroposki et al., DG power quality, protection and reliability case
studies report, National Renewable Energy Laboratory, Golden, CO,
NREL/SR-560-34635, 2003.
[40] C. Whitakar, J. Newmiller, M. Ropp, and B. Norris, Renewable system
interconnection study: distributed photovoltaic systems design and
technology requirements, Sandia National Laboratories, Livermore,
CA, SAND2008-0946, 2008.

1436

[41] Grid reliability and power quality impacts of distributed resources,. Palo Alto, CA, EPRI, 2003.
[42] M. Baran, Fault analysis on distribution feeders with distributed generator, IEEE Trans. Power Syst., vol. 20, no. 4, pp. 17571764, Nov.
2005.
[43] R. J. Bravo, R. Yinger, S. Robles, and W. Tamae, Solar PV inverter
testing for model validation, in Proc. IEEE Power Energy Soc. Gen.
Meet., 2011, pp. 17.
[44] P. M. Anderson, Power System Protection. New York: IEEE Press,
1999, pp. 220222.
[45] H. Li, F. Li, Y. Xu, D. T. Rizy, and J. D. Kueck, Adaptive voltage
control with distributed energy resources: Algorithm, theoritical analysis, simulation, ans field test verification, IEEE Trans. Power Syst.,
vol. 25, no. 3, pp. 16381647, Aug. 2010.
[46] M. Prodanovic, K. D. Brabandere, J. Van den Keybus, T. Green, and J.
Driesen, Harmonic and reactive power compensation as ancillary services in inverter-based distributed generation, IET Gener., Transm.,
Distrib., vol. 1, no. 3, pp. 432438, 2007.
[47] A. Bracale, R. Angelino, G. Carpinelli, M. Mangoni, and D. Proto,
Dispersed generation units providing system ancillary services in distribution networks by a centralised control, IET Renewable Power
Gener., vol. 5, no. 4, pp. 3113218, 2009.
[48] IEEE 1547-2003, IEEE Standard for Interconnecting Distributed Resources with Electric Power System, , 2003.
[49] C. Rahmann, H. J. Haubrich, A. Moser, R. Palma-Behnke, L. Vargas,
and M. B. C. Salles, Justified fault-ride through requirements for wind
turbines in power systems, IEEE Trans. Power Syst., vol. 26, no. 3, pp.
15551563, Aug. 2011.
[50] E.ON Netz grid codeHigh and extra high voltage, E.ON Netz
GmbH. Bayreuth, Germany [Online]. Available: http://www.eonnetz.com/Ressources/downloads/enenarhseng1.pdf
[51] Wind turbines connected to grids with voltages above 100 kVTechnical regulation for the properties and the regulation of wind turbines,
Elkraft System and Eltra Regulation, Draft version TF 3.2.5, 2004.
[52] E. Troester, New German grid codes for connecting PV systems to
the medium voltage power grid, in Proc. 2nd Int. Workshop Concentrating Photovoltaic Power Plant, 2009.
[53] Technical guidelineGenerating plants connected to the mediumvoltage network, Bundesverband der Energie-und Wasserwirtschaft
e.V.(BDEW), Berlin, Germany, 2008.
[54] C. A. Plet, M. Graovac, T. C. Green, and R. Iravani, Fault response
of grid-connected inverter dominated networks, in Proc. IEEE Power
Energy Soc. Gen. Meet., 2010, pp. 18.

IEEE TRANSACTIONS ON SMART GRID, VOL. 3, NO. 3, SEPTEMBER 2012

[55] W. H. Kersting, Radial distribution test feeders, in Proc. IEEE Power


Eng. Soc. Winter Meet., 2001, pp. 908912.
[56] Network protection and automation guide, AREVA T&D Automation,1st ed. , 2002.

Hesam Yazdanpanahi (S11) received the B.Sc. and M.Sc. degrees in electrical engineering from Amirkabir University of Technology, Tehran, Iran, in
2007 and 2009, respectively. He is currently working toward the Ph.D. degree
in electrical and computer engineering at the University of Alberta, Edmonton,
Canada.
His research interests are distributed generation (DG) and power system
protection.

Yun Wei Li (S04M05SM11) received the B.Sc. in Engineering degree in


electrical engineering from Tianjin University, Tianjin, China, in 2002, and the
Ph.D. degree from Nanyang Technological University, Singapore, in 2006.
In 2005, he was a Visiting Scholar with Aalborg University, Denmark, where
he worked on the medium voltage dynamic voltage restorer (DVR) system.
From 2006 to 2007, he was a Postdoctoral Research Fellow at Ryerson University, Canada, working on the high power converter and electric drives. In 2007,
he also worked at Rockwell Automation Canada and was responsible for the development of power factor compensation strategies for induction motor drives.
Since 2007, he has been an Assistant Professor with the Department of Electrical and Computer Engineering, University of Alberta, Edmonton, Canada.
His research interests include distributed generation, microgrid, renewable energy, power quality, high power converters and electric motor drives.
Dr. Li serves as an Associate Editor for the IEEE TRANSACTIONS ON
INDUSTRIAL ELECTRONICS and a Guest Editor for the IEEE TRANSACTIONS
ON INDUSTRIAL ELECTRONICS Special Session on Distributed Generation and
Microgrids.

Wilsun Xu (F05) received the Ph.D. degree from the University of British Columbia, Vancouver, Canada, in 1989.
From 1989 to 1996, he was an Electrical Engineer with BC Hydro, Vancouver,
Canada. Currently, he is with the University of Alberta, Edmonton, Canada as a
Research Chair Professor. His current research interests are power quality, information extraction from power disturbances, and power system measurements.

You might also like