Вы находитесь на странице: 1из 9

110

IEEE TRANSACTIONS ON ENERGY CONVERSION, VOL. 29, NO. 1, MARCH 2014

New Generic Model of DFG-Based Wind Turbines


for RMS-Type Simulation
Jens Fortmann, Member, IEEE, Stephan Engelhardt, Member, IEEE, Jorg Kretschmann, Member, IEEE,
Christian Feltes, and Istvan Erlich, Senior Member, IEEE

AbstractNew requirements for the validation of simulation


models based on measurements in many grid codes show that existing generic approaches for generator and converter models of
doubly fed generator systems (DFG) may not be accurate enough.
The authors show that by applying a detailed analysis of the generator equations and the converter control design, a reduction of
the model complexity is possible while maintaining a high level of
accuracy. The generator model presented in this paper allows an
improved representation of the stationary and dynamic response
of wind turbines equipped with DFG systems especially during
grid faults and during voltage recovery. The model is designed to
represent modern DFG systems independently of vendor specific
hardware and software. The results of simulations are compared
to measurements of a voltage dip involving wind turbines. The generator model has been proposed as extension to the WECC/IEEE
generator models and has been accepted as reference for IEC TC88
working group 27 (standard IEC 61400-27-1) on modeling and
model validation of wind turbines.
Index TermsControl, IEC 61400-27, measurement, modeling,
standards, wind energy.

I. INTRODUCTION
ENERIC models for DFG-based wind turbines (WTs)
have been proposed by WECC [1] (DFG is described
as type 3 by WECC). Simulation results of the DFG were
compared to simulations with a more detailed model in [2].
Those models were so far mainly used in the USA, and the
models were able to represent WTs with respect to common
U.S. grid code requirements.
European grid codes often have specific requirements on reactive power delivery during and following grid faults [3][5].
New regulations in Germany [6], [7] and Spain [8] require the
validation of WT simulation models with measurements of voltage dips. But this is beyond the capabilities of the existing
generic models.
A very detailed approach to modeling DFG systems had been
presented in [9], but the level of detail requires both a large

Manuscript received February 10, 2013; revised May 26, 2013 and August
20, 2013; accepted September 14, 2013. Date of publication November 7, 2013;
date of current version February 14, 2014. Paper no. TEC-00081-2013.
J. Fortmann is with the REpower Systems AG, 24783 Osterronfeld, Germany
(e-mail: j.fortmann@repower.de).
S. Engelhardt and J. Kretschmann are with Woodward SEG GmbH &
Co, 47906 Kempen, Germany (e-mail: stephan.engelhardt@woodward.com;
joerg.kretschmann@woodward.com).
C. Feltes and I. Erlich are with the University Duisburg-Essen, 47057 Duisburg, Germany (e-mail: christian.feltes@uni-due.de; istvan.erlich@uni-due.de).
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/TEC.2013.2287251

Fig. 1.

Wind turbine model structure.

number of parameters and integration step times smaller than


1 ms. In this paper, a new approach is presented that tries to
come close to the accuracy of detailed models while keeping a
simple structure and a very limited number of parameters as it
is common for generic models. This is achieved on the basis of
a detailed analysis of the generator equations and the converter
control design.
The model can be used both for large-scale stability-type rootmean-square (rms) simulation and with an integration time step
of 110 ms. Time constants of less than 5 ms have been eliminated by aggregating the generator and controller equations.
The generator and converter model as part of the WT model is
shown in Fig. 1. The model is intended to represent the aspects
of a generator as well as of a converter that have relevance for
large scale stability type rms simulation. Some of the key issues
it has to reproduce properly are:
1) response to voltage amplitude and phase angle changes;
2) response to active and reactive power set point changes.
According to the objective, transient effects should not be
modeled in detail. During transient periods it is sufficient to
represent fast changing variables by their average. The model is
supposed to have a simple structure and should only use a limited number of parameters. On the other side, the physical background of parameters should remain visible, so that the value of
parameters can be traced back to a physical or control quantity.
The model should be simple enough not to disclose manufacturer specific know-how while still being accurate enough to
allow a certification required by several grid codes.
II. MODEL OF THE DFG-BASED WT
The DFG is the most commonly used device for wind power
generation. The stator is connected to the grid, the rotor terminals are fed in normal operation with symmetrical three-phase

0885-8969 2013 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See http://www.ieee.org/publications standards/publications/rights/index.html for more information.

FORTMANN et al.: NEW GENERIC MODEL OF DFG-BASED WIND TURBINES FOR RMS-TYPE SIMULATION

111

For the modeling approach selected, the dc-link voltage is


therefore assumed to remain close to its nominal value and
therefore does not need to be modeled explicitly.
B. Basic DFG Equations
The voltage and flux equations of the DFG in a dq coordinate
system rotating synchronously with the grid frequency 0 are
given by the following equations:
v S = rS iS +
v R = rR iR +

d S
dt
d R

+ j0 S

dt
S = lS iS + lm iR
Fig. 2.

A. Limitations
1) Limitations of DFG Designs Without DC-Link Energy
Absorber: In older DFG designs, protection against overcurrents and undesirably high dc voltage was provided by the crowbar (CR) placed on the rotor side of the machine side converter
(MSC). Very high rotor current following voltage dips could
cause the dc voltage to exceed the upper threshold allowed. The
CR thyristor switches were then activated and the rotor terminals
short-circuited through the CR resistance [9], [10]. As a result,
the DFG then is operated as a conventional slip-ring induction
machine without any control capability left for the MSC.
The reaction to grid faults of all these designs is highly nonlinear and usually requires detailed electromagnetic transient
(EMT)models to exactly represent the dc-link that leads to
the triggering of the crowbar. In [9], a modeling approach was
presented that allows a detailed simulation even with rms model
but very small simulation time steps.
2) Improved Protection of Modern DFG by Using DC-Link
Chopper: New German grid code requirements [4] demand the
provision of 0.9 p.u. reactive current within 30 ms of a voltage
drop below 0.5 p.u. rated voltage. To achieve this, newer DFG
designs use a higher rating of the MSC and passive or active
dc-link elements. One common approach is using an IGBTswitched resistor connected to the dc-link, referred to as chopper (CH), which limits the voltage in case the power fed into the
dc-link is higher than the power the line side converter (LSC)
can feed into the grid. As a result, an activation of the crowbar during and following grid faults is no longer necessary and
MSC and LSC remain controllable during the entire fault period. Crowbar-based converter protection may still be available
as backup protection, but it is usually not activated during grid
faults.

(2)
(3)

R = lm iS + lR iR

Key components of DFG system.

voltage of variable frequency and amplitude. This voltage is


provided by a voltage source converter usually equipped with
IGBT-based power electronics circuits. The basic structure is
shown in Fig. 2.

+ j (0 R ) R

(1)

(4)

with lS = lm + l S , lR = lm + l R and the complex space


vectors S = S d + jS q , R = R d + jR q , v S = vS d +
jvS q , v R = vR d + jvR q , iS = iS d + jiS q , and iR = iR d +
jiR q .
In this paper, the authors are using normalized p.u. quantities
referred to WT nominal power, voltage and current, respectively.
The synchronous speed is characterized by 0 , which is 1.0 p.u.
The consumer-oriented sign convention is applied for all equations and diagrams where consumed active power and inductive
reactive power are considered to be positive.
C. DFG Model Representation
For stability type simulation, similarly to synchronous machines the derivative term in the stator voltage equation is set
to zero (d S /dt = 0) [11]. Then, from (1) after some algebraic
manipulation and by using (3) and (4) follows:
v S = z  iS + v 

(5)

where the internal transient impedance is defined as


z  = rS + j0 l

(6)

with the transient inductance


l  = lS

2
lm
lR

(7)

and the voltage behind the transient reactance


v  = j0

lm
= j0 kR R
lR R

(8)

where the coupling factor


kR =

lm
lR

(9)

is additionally introduced. Fig. 3 shows the corresponding


Thevenin voltage source model of the reduced order DFG
model.
The voltage source v  can be calculated by solving the rotor
flux differential (10).
d R
dt

rR
j (0 R ) R + kR rR iS + v R . (10)
lR R

112

Fig. 3.

IEEE TRANSACTIONS ON ENERGY CONVERSION, VOL. 29, NO. 1, MARCH 2014

Thevenin equivalent for coupling of DFG with the grid.

Fig. 5.

Fig. 4.

by the MSC is determined by extending the feed-forward term


(15) by the output of an additional current controller v C
R ,ref

Block diagram of the DFG reduced order model.

It follows from (2) by replacing the rotor current iR with


iR =

R
lR

kR iS

(11)

which is derived from (4). For the stator current, the following
expression can be deduced from (5)
iS =

vS
v

.
z
z

Model of MSC and corresponding control.

(12)

The complete DFG model characterized by (5)(12) can be


described by a control schema shown in Fig. 4 where the following substitutions are introduced:
lR
TR =
rR

(13)

= (0 R ) .

(14)

and

D. MSC Control Representation


A common approach for the control of DFG systems is the
field oriented control as described in [9], [12], and [15]. In
this approach, an outer power control loop provides the current reference for the underlying active and reactive current
controllers. The current controller is developed based on (10)
by assuming steady-state conditions, i.e., d R /dt = 0. Then,
the corresponding rotor voltage can be directly calculated by
rearranging (10)


1
vD
=
+
j
(

)
R kR rR iS .
(15)
0
R
R ,ref
TR
To underline that the required rotor voltage v R ,ref may differ slightly from that calculated by (15), the superscript D is
introduced indicating the direct steady-state contribution. The
control effect of v D
R ,ref is similar to that of a feed-forward controller as will be demonstrated in Fig. 5. However, in real applications, it is usually calculated by using the instant values of
rotor flux and stator current instead of the corresponding reference values. The entire rotor voltage reference to be generated

C
v R ,ref = v D
R ,ref + v R ,ref .

(16)

The task of this controller with usually PI characteristic is,


on the one hand, to compensate for parameter and measurement
uncertainties and, on the other hand, to improve the dynamic
control behavior


1
vC
=
k
(17)
1
+
iR
C ,R
R ,ref
sTC
With kC ,R the proportional gain of the rotor current control and
TC the integral time constant of the current control and
iR = iR ,ref iR .

(18)

As the direct feed-forward control contribution results in decoupling of the control channels (note that complex quantities
are used for deriving the model), the additional current controllers according to (17) will control the real and imaginary
rotor current components independently.
For multimegawatt turbines, the stator resistance is very
small, and thus negligible for the controller design. Then, the
relation between stator and rotor currents under steady-state
conditions can be written by using (1) and (3) as
iR =

lS
v
i j S.
lm S
lm

(19)

Under the simplified conditions that the stator voltage is considered to be constant, the rotor current deviation can be calculated as

lS 
iS,ref iS
(20)
iR =
lm
where the stator reference current iS ,ref is introduced. The error
resulting from the assumption of a fixed stator voltage will
affect only the reactive current. Besides, it can be compensated
by slightly modified parameter of the PI controller.
For the rotor current controller then can be written




1
C
(21)
v R ,ref = kC ,S 1 +
iS,ref iS
sTC

FORTMANN et al.: NEW GENERIC MODEL OF DFG-BASED WIND TURBINES FOR RMS-TYPE SIMULATION

Fig. 7.

Fig. 6.

113

Aggregated model of DFG and MSC.

Model of DFG and MSC and possible simplification.

with
kC ,S =

lS
kC ,R .
lm

(22)

The control schema of the MSC based on the equations discussed earlier is shown in Fig. 5. The converter itself is represented by a first-order delay block.
The rotor flux required in Fig. 5 can be calculated from (4).
However, due to parameter and measurement uncertainties, it
may differ slightly from the real rotor flux. Therefore, it is
denoted as R . In practical implementations usually observer
based methods are used to determine the rotor flux providing more accurate results [13]. In the simulation model, however, parameter and measurement tolerances do not need to be
considered.
E. Combined DFG and MSC Model
For better understanding of the following steps the DFG and
the MSC models are drawn in a single diagram shown in Fig. 6.
Taking into consideration that the time constant of the converter TM SC is small in comparison with the delay caused by
the generator the block representing the converter can be eliminated. With the assumption R R it becomes obvious that
the feedback through (1/TR + j) in the DFG model and the
corresponding term in the controller compensate each other and
can be deleted. The same applies to the terms kR rR iS . This
also becomes visible if vr in (10) is replaced by (15) and (21).
A further simplification is possible by incorporating the term
j0 kR /z  into the PI controller gain. With the assumption
rs  j0 l and thus z  jx = j0 l the control gain becomes


lS
j0 kR
kC ,S kR
=
kC ,R . (23)
kC = kC ,S
=
j0 l
l
lR l 
The simplifications steps describe earlier lead to the aggregated MSC and DFG model shown in Fig. 7.

Fig. 8. Aggregated model of DFG and MSC taking into consideration the
reference frames.

to separate between active and reactive current controllers. This


is achieved by defining the d-axis along the terminal voltage.
In such a coordinate system the active current will always correspond with the real part of the stator current and the reactive
current with the reverse (negative sign) of the imaginary component. However, it should be mentioned that many authors use
flux orientation that results in active current in the q-axis and
reactive current in the d-axis. In this paper, however, terminal
voltage orientation will be preferred. The DFG and MSC model
by taking into consideration the required coordinate transformations to establish active and reactive current control is shown
in Fig. 8.
The angle v represents the position of the terminal voltage
in grid synchronous coordinates and is determined usually by
a so called phase-locked-loop (PLL) unit. In a simulation environment, this angle is obtained by solving the grid equations.
The DFG, due to the direct connection to the grid, remains in
grid coordinates.
The current iPQ
S ,ref is defined as
Q
P
iPQ
S ,ref = iS ,ref jiS ,ref

(24)

with the reference values for stator active and reactive currents
pS ,ref
|v S |
qS ,ref
.
=
|v S |

iPS ,ref =

(25)

iQ
S ,ref

(26)

F. Selection of Reference Frame


For the derivation of the DFG equations in Section II only
the rotational speed of the coordinate system has been defined
by assuming grid synchronous frequency 0 but not the right
position of the dq axes. In a WT control system, it makes sense

G. Incorporation of the LSC Model


The basic design of the LSC is shown in Fig. 9. For steadystate operation and if losses are neglected, the active power of
the LSC fed into the grid is defined by the stator power pS and

114

Fig. 9.

IEEE TRANSACTIONS ON ENERGY CONVERSION, VOL. 29, NO. 1, MARCH 2014

LSC design.

the rotor speed R [12], [13], [15]


pLSC =

R 0
pS .
0

(27)

The total active power of turbine pW T is the sum of the stator


power pS and power via the LSC pLSC .
pW T = pS +

R 0
R
pS =
pS .
0
0

(28)

The active power control of the LSC usually bases on the


control of the dc-link voltage [13]. The response time of the
LSC can be further reduced by adding the MSC power, which
corresponds with the rotor power, as reference to the LSC control
[13]. The LSC reactive power can be controlled independently
within the voltage and current limits of the converter. In general
the LSC response is faster than that of the MSC supplying
through the DFG to the grid [13]. Therefore, in a simplified
generic model the contribution of LSC can be considered by the
steady-state equations without modeling the LSC controller in
detail.
In controller design, a compensation of the speed dependency
of turbine power in (28) as shown in (29) is commonly used for
deriving the turbine currents from the stator currents (see [13,
p. 60])
R P
Q
P
(i jiQ
(29)
iPQ
W T = iW T jiW T
S ).
0 S
For the active current, (29) follows directly from (28). For
the reactive current, as described earlier using reactive power
from the LSC offers an additional degree of freedom to compensate the impact of the rotor speed. As a result, the WT control
becomes independent of the operating speed under normal operating conditions.
Under transient conditions following fast voltage amplitude
or voltage phase angle changes, the amplitude or the rate of
change of currents from the generator may not immediately
be fully compensated by the converter control. The impact of
transient voltage changes is represented by the impedance x of
the generator. At rated rotor speed R ,n , using (28) the generator
currents are actually lower than the turbine currents by the factor
0 /R ,n . It is assumed that R equals nominal rotor speed R ,n
(R = R ,n ), which is the case in modern WT already at about
approximately 0.4 p.u. active power (see [14, p. 531]) due to
increased rotor diameters while at the same time the tip speed
of the blades is kept constant in order to keep acoustic impact
of turbine at its present level. As a result, this assumption is true
for most of the operating range but slightly underestimates to
contribution of x to the transient currents around synchronous
speed.

Fig. 10.

Aggregated WT model.

Fig. 11.

Coupling of WT model with the grid via current source model.

By using the speed compensation of (29) and splitting the


model in Fig. 8 in real and imaginary components for better
clarity the WT model including the DFG, MSC, and LSC is
shown in Fig. 10.
A new equivalent reactance can be introduced as follows:
0
.
(30)
xE = x
R ,n
In stability kind of power system simulation software, the grid
is usually described by complex algebraic equations. Therefore,
following grid faults the voltage magnitude and phase angle
can change immediately. To improve the interaction of the WT
model with the grid and thus improve the convergence behavior
of the simulation, it is recommended to incorporate the term
stator voltage divided by the transient reactance into the grid
equations. The necessary modification is shown in Fig. 11. For
a consistent notation the voltage has been changed to v W T from
v S even though both are the same.

FORTMANN et al.: NEW GENERIC MODEL OF DFG-BASED WIND TURBINES FOR RMS-TYPE SIMULATION

Fig. 12.

Most simplified DFG model.

Fig. 14.

Fig. 13.

115

WT speed controller.

H. Most Simplified Aggregated DFG and MSC/LSC Model


A further simplification of the DFG model is possible if the
impact of the transient reactance is neglected. The simplest
model can be derived by representing the behavior of the WT
by a first-order delay only [16], [17]. The corresponding model
is shown in Fig. 12.
The model is based on the assumption that the WT response
is faster than the response of other power system components
and the transient period not correctly represented in this model
will not significantly affect the phenomena under investigation.
But as will be shown later, neglecting the transient impedance
will lead to voltage spikes during fault clearance and does not
represent effects related to voltage phase angle changes. The
model has only one parameter that is the equivalent time constant
TW T . However, to calculate the correct reference values the
models described in the next section have to be used, as is the
case in the other more detailed models too.
I. Calculation of Active and Reactive Current References
The required active current is derived from the active power
the WT has to supply. This, on the other hand, is the output of
the speed controller, i.e., the reference is provided as a function of the measured WT power output. An additional filter
ensures that fast power changes will not result in change of the
speed reference. Manufacturer specific control is implemented
for proper coordination with the pitch controller (see also [14,
p 572ff], [16, p 36ff], [18, p. 480ff]) and for handling the fault
ride-through period and the subsequent recovery phase. Most of
the WT are also equipped with a shaft damping controller. The
corresponding control schema is depicted in Fig. 13.

Cascading reactive power and voltage controller.

The generated reactive power can be defined within the technical limitations of the WT [18]. Common method to describe
the reactive power demand is specifying the power factor or
a reactive power/voltage characteristic. Both are illustrated in
Fig. 14. In steady-state operation, the required response characteristic of the reactive power controller is usually not demanding. Response times in the range of several seconds up to one
minute are suitable as the control has to compensate the slow
load flow changes within the grid only. Following sudden grid
voltage drops, however, the WT has to inject reactive current
very fast. In [4], 30 ms rise and 60 ms settling time are required. A cascaded control schema that meets both (different)
time requirements is shown in Fig. 14.
The inner voltage controller reacts fast whereas the outer
reactive power controller is slow. Due to its PI characteristic
the reactive power controller will determine the WT behavior in
steady-state. However, throughout short fault period the output
of this controller does not change considerably. Therefore, the
fast voltage controller will dominate the WT behavior during
the fault.
The active and reactive current references must be limited before passing on the reference to the current controller. Assuming
the maximum WT current allowed is iW T,m ax the limitations
are depending on the priority settings as follows.
Active current priority
iPW T,ref = iPWT,ref .

iQ
=
(iW T,m ax )2 (iPWT,ref )2
W T,ref

(31)

Reactive current priority


Q
iQ
W T,ref = iW T,ref .


2
iPW T,ref = (iW T,m ax )2 (iQ
W T,ref )

(32)

The priorities basically are decided based on the voltage


level. In normal operation the active current is prioritized. However, following grid faults below 80%90% voltage forced reactive current for voltage support is required in several grid
codes [3][5]. During this stage, the active current supply is less
important and can be reduced in favor of the reactive current.
Additional limitations will usually be applied depending on the
grid voltage and frequency. In normal operation, steady-state
limits apply, and at lower voltages temporary higher currents
are usually allowed during a fault. In weak grids for stability

116

Fig. 15.

IEEE TRANSACTIONS ON ENERGY CONVERSION, VOL. 29, NO. 1, MARCH 2014

Identification of generic model parameters.

reason, e.g., it may be necessary to reduce the active current beyond the value given in (32) during the low voltage period [19].
A common implementation of such a functionality is a separate lookup table or an analytical relationship for active and
reactive current as function of voltage. Also following voltage
recovery it makes sense to increase the active current with a
limited ramp rate. Common implementations are either a rate
limiter on active power or active current or torque or a resetting
of the torque controller during faults as shown in Fig. 13. It
should be mentioned that the maximum possible current is not
fixed but variable depending on the instantaneous temperature
of the power electronic components. How and to what extent
such limitations are implemented is specific to manufacturers
and cannot be described in a general model.

III. CALCULATION OF MODEL PARAMETERS


The WT models described in the previous sections are based
on simplifications and aggregations. That means the corresponding generator parameter do not represent necessarily real WT
parameters but rather approximations. Converter control parameters and limits will usually be deduced from parameters of the
real system. Generator parameters (kS , TS , xE ) can be deduced
from physical parameters, but an improved model accuracy can
often be achieved by optimizing these values using identification
techniques. The identification procedure is illustrated in Fig. 15.
In this study, the heuristic optimization algorithm called mean
variance mapping optimization (MVMO) has been used [20].
As reference for comparison with the generic model measurements or accurate simulation results from models validated by
measurements can be used. Those results should cover typical
contingencies for which the model is intended to be used.
It is also recommended to carry out the identification for several scenarios simultaneously. This will guarantee that the model
parameters are not only tuned for a single case providing good
accuracy but also for other scenarios that are poorly represented.

Fig. 16. Measurement (solid line) and simulation dashed line) of voltage,
reactive and active power during a voltage dip down to 20% rated voltage for
wind turbine with DFG operating at rated power using the proposed aggregated
model shown in Fig. 10.

IV. VALIDATION
A measurement setup according to [6] has been used to
compare the results of simulations with measurements for a
2-MW DFG-based WT. The producer oriented sign convention
is applied for all validation figures, generator active power and
overexcited reactive power are considered to be positive. The
main parameters used for the simulation in this chapter are listed
in the appendix.
A. Measurements and Simulations
For the comparison, three-phase measurement data is decomposed into positive and negative sequence components according to [23]. The output of the simulation is filtered with a 50 Hz
average filter in order to have an equivalent to the positive and
negative sequence values of measurements [6].
Measurements of a voltage dip down to 20% of the rated
voltage of a DFG are compared with simulations in Figs. 16 and
17 for a 2 MW WT. The voltage reference value for the inverter
(vW T,ref , see Fig. 14) of measurement and simulation has been
kept constant. The active power reference value for the inverter
(pW T,ref ) is calculated by the turbine control (see Fig. 13). This
value has been recorded during measurement and is used as
input for the simulation model. This is necessary to evaluate the
accuracy of the generator and inverter model only and helps to
limit uncertainties on the reference value arising from both the
aerodynamic, mechanical and control models of the turbine and
uncertainties of the wind speed measurement.

FORTMANN et al.: NEW GENERIC MODEL OF DFG-BASED WIND TURBINES FOR RMS-TYPE SIMULATION

117

V. CONCLUSION
A new generic generator model for the DFG system for rms
modeling is proposed. The model is derived from the physical
equations of the generator, the converter hardware and the converter control system. Compared to existing generic modeling
approaches, the proposed model offers higher accuracy and a
clear physical deduction of the topology and of all parameters.
Models for active power control, reactive power control and
current limitation are proposed that allow a generic parameterization independent of the manufacturer-specific implementation. The requirements for implementing the model in grid
simulation software are explained.
The proposed model shows very good correlation with measurements of DFG WTs while still requiring only a very limited
number of additional parameters compared to existing, simpler
modeling approaches.
The generator and inverter model has been accepted as reference model for the generic type 3 (DFG) models in the new
standard on WT modeling IEC 61400-27-1 [24].
APPENDIX
Fig. 17. Comparison of measurement of DFG (solid line) and simulation using
the proposed generator model (dashed line) and using the simplified first order
lag model (dash-dotted line) for active power and reactive power during voltage
dip down to 20%% rated voltage and voltage recovery.

The simulation shows a very good correlation of both active


and reactive power before, during and after the voltage dip. The
small oscillation of the measured current at t = 4.2 s is due to
the closing of the switch bypassing the series impedance of the
test equipment (see [6]) and the resulting change of the voltage phase angle. The simulation is done using voltage feedback
from the measurement to avoid associating uncertainties of the
grid model to the inverter and generator model. In this way,
physical phenomena like the transformer inrush visible in measurements during voltage recovery at t = 3.2 s are also taken in
consideration.
The details of voltage drop and voltage recovery from Fig. 16
are shown in Fig. 17. Measurement and simulation using the
proposed aggregated model shown in Fig. 10 are compared
with the simplified model according to Fig. 12.
The figure shows the reactive power of a DFG-based turbine
at the beginning and at the end of a voltage dip. As can be seen,
the aggregated model shows a far better representation, during
the first 50 ms following the voltage drop and following the
voltage recovery.
Especially the reactive power spike following the voltage
recovery of the simplified model can lead to voltage spikes
that can cause stability problems during simulations by triggering protection equipment action. The simplified model in
Fig. 12 proposed by [16] is identical to the model used for full
converter systems (see [16, p. 3-1) wherecontrary to DFG
systemssuch spikes actually occur [21]. By contrast, the proposed model (dashed line) shows a very good correlation with
the reactive power change directly following the voltage recovery and actually limits the voltage increase to the same level as
the measurement.

The following blocks with their main parameters and input


values have been used for validation:
Voltage controller (see Fig. 14): kV = 4, vW T,ref = 0,
vW T,ref0 = 1.
Current limitation: iW T,m ax = 1.4 p.u., iW T,m ax,P = 1.1 p.u,
iW T,m ax,Q = 1,
Active power reference (see Fig. 13): pW T,ref from measurement.
Generator model (see Figs. 10 and 11): xE = 0.48 p.u., kC =
90 p.u., Ti = 0.2 s, R ,n = 1.2 p.u. The value of x from the
datasheet was 0.36 p.u., the calculated value of xE = x 1.2 =
0.43 p.u..
Active power controller parameters (For information only,
not used for validation): TW T = 0.02 s, T ref ,n = 5 s, kS =
7 p.u., TS = 0.166 s, tW T,m ax = 1 p.u., and Oset = 0.02 p.u.
REFERENCES
[1] Generic Type-3 Wind Turbine-Generator Model for Grid Studies, WECC
Wind Generator Modeling Group, Version 1.1, Irwindale, CA,USA, Sep.
14, 2006.
[2] W. Price and J. Sanchez-Gasca, Simplified wind turbine generator aerodynamic models for transient stability studies, in Proc. IEEE PES 2006
Power Syst. Conf. Expo. (PSCE), Atlanta, GA, Oct. 29/Nov 1.
[3] E.ON Netz GmbH. (2006). Netzanschlussregelnhoch-und hochstspannung [Online]. Available: http://www.eon-netz.com/pages/ehn_de/
Veroeffentlichungen/Netzanschluss/Netzanschlussregeln/ENENARHS20
06de.pdf
[4] BMU. (2009). Verordnung zu Systemdienstleistungen durch Windenergieanlagen (SystemdienstleistungsverordnungSDLWindV) [Online]. Available: http://www.gesetze-im-internet.de/bundesrecht/sdlwindv/gesamt.
pdf
[5] (2006). REERequisitos de respuesta frente a huecos de tension
de las instalaciones de produccion de regimen especial, PO 12.3
[Online]. Available: http://www.ree.es/operacion/pdf/po/PO_resol_12.3_
Respuesta_huecos_eolica.pdf
[6] FGW: Technical Guidelines for Power Generating Units. Part 4 Demands
on Modeling and Validating Simulation Models of the Electrical Characteristics of Power generation Units and Systems, Revision 6, May 1,
2013.

118

[7] FGW: Technical Guidelines for Power Generating Units. Part 8. Certification of the Electrical Characteristics of Power Generating Units and
Farms in the Medium-, High- and Highest-voltage Grids 6, May 1, 2013.
[8] Procedure for Verification Validation and Certification of the Requirements of the PO 12.3 on the Response of Wind Farm in the Event of
Voltage Dips, AEE [Online]. Available: http://www.aeeolica.org/uploads/
documents/1306-pvvc-n9-english.pdf
[9] I. Erlich, J. Kretschmann, J. Fortmann, S. Engelhardt, and H. Wrede,
Modeling of wind turbines based on doubly-fed induction generators for
power system stability studies, IEEE Trans. Power Syst., vol. 22, no. 3,
pp. 909919, Aug. 2007.
[10] I. Martinez de Alegira, J. L. Villate, J. Andreau, I. Gabilola, and P. Ibanez,
Grid connection of doubly fed induction generator wind turbines: A
survey presented at Eur. Wind Energy Conf. 2004 [Online]. Available:
http://www.2004ewec.info
[11] P. Kundur, Power System Stability and Control. New York, NY, USA:
McGraw-Hill, 1994. ISBN 007035958-X.
[12] G. Michalke, Variable speed wind turbinesModelling, control, and
impact on power systems, Ph.D. dissertation, TU Darmstadt, Germany,
2008.
[13] S. Engelhardt, Direkte Leistungsregelung einer Windenergieanlage mit
doppelt gespeister Asynchronmaschine, Ph.D dissertation, Universitat
Duisburg-Essen, Germany, 2011.
[14] T. Ackermann, Wind Power in Power Systems. Hoboken, NJ, USA: Wiley, 2005.
[15] D. Arsudis, Doppeltgespeister Drehstromgenerator mit Spannungszwischenkreis-Umrichter im Rotorkreis fur Windkraftanlagen, Ph.D. dissertation, TU Braunschweig, Germany, 1989.
[16] P. Pourbeik. (2013). Proposed changes to the WECC WT3 generic model
for type 3 wind turbine generators. Electr. Power Res. Inst. Online. Available: http://www.wecc.biz/library/WECC%20Documents/Documents%
20for%20Generators/WECC%20Type%203%20Wind%20Turbine%20
Generator%20Model%20-%20Phase%20II%20012313.pdf
[17] P. Pourbeik. (2012). Proposed changes to the WECC WT4 generic model
for type 4 wind turbine generators. Electric Power Res. Inst. [Online].
Available: http://www.wecc.biz/committees/StandingCommittees/PCC/
TSS/MVWG/REMTF/Shared%20Documents/Report_on_WT4_Model_
Description_PP070312.pdf
[18] T. Burton, D. Sharpe, N. Jenkins, and E. Bossanyi, Wind Energy Handbook. Hoboken, NJ, USA: Wiley, 2001.
[19] I. Erlich, F. Shewarega, S. Engelhardt, J. Kretschmann, J. Fortmann, and
F. Koch, Effect of wind turbine output current during faults on grid
voltage and transients stability of wind parks, presented at 2009 IEEE
Power Energy Soc. General Meeting, Calgary, AB, Canada.
[20] J. L. Rueda and I. Erlich, Optimal dispatch of reactive power sources
by using MVMOS optimization, presented at 2013 IEEE Symp. Series
Comput. Intell. (IEEE SSCI 2013), Singapore, 1619, Apr.
[21] S. Ledwon, F. Kalverkamp, J. Langstadtler, and B. Schowe-von-derBrelie, Dynamic Voltage support of decentralized Power Generators,
presented at 11th Int. Workshop Large-Scale Integr. Wind Power Power
Syst. Well Transmiss. Netw. Offshore Wind Farms, Lisbon, Portugal,
2012.
[22] S. Engelhardt, I. Erlich, C. Feltes, J. Kretschmann, and F. Shewarega,
Reactive power capability of wind turbines based on doubly fed induction
generators, IEEE Trans. Energy Convers., vol. PP, no. 99, pp. 19, Oct.
2010.
[23] Wind Turbine Generator SystemsPart 21: Measurement and Assessment
of Power Quality Characteristics of Grid Connected Wind Turbines, IEC:
IEC 61400-21 ed. 2, Aug. 13, 2008.
[24] Wind TurbinesPart 27: Electrical Simulation Models for Wind Power
GenerationSection 1, CDV 2013-07-31, IEC: IEC 61400-27-1 ed. 1,
Nov. 11, 2013.
Jens Fortmann (M05) was born in 1966. He received the Dipl.-Ing. degree in electrical engineering
from the Technical University, Berlin, Germany, in
1996.
From 1995 to 2002, he was involved in the simulation of the electrical system and the control design of variable speed wind turbines at the different
wind turbine manufacturers. Since 2002, he has been
with REpower Systems AG, Osterronfeld, Germany,
where he is currently a Team Leader of model and
system development. He is the Head of the FGW
Working Group that specifies the modeling and model validation guideline
TR4.

IEEE TRANSACTIONS ON ENERGY CONVERSION, VOL. 29, NO. 1, MARCH 2014

Stephan Engelhardt (M05) was born in 1967. He


received the Dipl.-Ing. degree in electrical engineering from the University Hannover, Germany, in 1997.
Since 1997, he has been with Woodward SEG
GmbH & Co. KG, Kempen, Germany, where he is
currently the Head of the Group Converter Technology and responsible for system designs and simulations, control strategies and patents.

Jorg Kretschmann (M05) was born in 1958. He received the Dipl.-Ing. degree in electrical engineering
from the Technical University Berlin, Germany, in
1986.
From 1986 to 1988, he was with the Engineering Department, AEG-Kanis, Essen, manufacturing
of synchronous generators up to 200 MVA. Since
1988, he has been with Woodward SEG GmbH &
Co. KG, Kempen, Germany, as a Designing Engineer for speed-variable applications: uninterruptible power supply, shaft alternators, DFIG for wind
turbines.

Christian Feltes was born in 1979. He received the


Dipl.-Ing. degree in electrical engineering in 2005
from the University of Duisburg-Essen, Essen, Germany, where he is currently working toward the
Ph.D. degree in the Department of Electrical Power
Systems.
His current research interests include wind energy
generation, control, integration, and dynamic interaction with electrical grid.

Istvan Erlich (SM08) was born in 1953. He received the Dipl.-Ing. degree in electrical engineering
and the Ph.D. degree from the University of Dresden,
Dresden, Germany, in 1976 and 1983, respectively.
He was involved in different fields of power engineering in Hungary, Berlin and Dresden, Germany.
Since 1998, he has been a Professor and the Head
of the Institute of Electrical Power Systems, University Duisburg-Essen, Germany. His current research
interests include power system stability and control,
modeling and simulation of power system dynamics,
including intelligent system applications, smart grids, and renewable energy
sources.
Dr. Erlich is a member of VDE. He is also chairing the IFAC Technical
Committee 6.3 on Power and Energy Systems and the German Chapter of IEEE
PES.

Вам также может понравиться