You are on page 1of 23

2776

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 7, JULY 2013

Overview of Control Systems for the Operation


of DFIGs in Wind Energy Applications
Roberto Crdenas, Senior Member, IEEE, Rubn Pea, Member, IEEE,
Salvador Alepuz, Senior Member, IEEE, and Greg Asher, Fellow, IEEE

AbstractDoubly fed induction generators (DFIGs), often organized in wind parks, are the most important generators used
for variable-speed wind energy generation. This paper reviews the
control systems for the operation of DFIGs and brushless DFIGs
in wind energy applications. Control systems for stand-alone
operation, connection to balanced or unbalanced grids, sensorless control, and frequency support from DFIGs and low-voltage
ride-through issues are discussed.
Index TermsControl strategies, crowbar, doubly fed induction
generator (DFIG), low-voltage ride through (LVRT), reactive support, robust controller, voltage unbalance, wind turbine.

v1s
v2s
v0
vr
vf
sl
s
r
s

Positive sequence of the stator voltage.


Negative sequence of the stator voltage.
Zero sequence of the stator voltage.
Rotor voltage vector.
GSC voltage vector.
Slip angle.
Position of the stator-flux vector.
Rotor position.
Stator time constant.
I. I NTRODUCTION

N OMENCLATURE
ims
is
ir
if
J
ktransf
Ls
Lr
L0
e
r
sl
p
s
r
Rs
Rr
s
Te
vs

Magnetizing current vector.


Stator current vector.
Rotor current vector.
Grid-side converter (GSC) current vector.
Rotor inertia
Constant for the abc-to-dq transformation.
Stator inductance.
Rotor inductance.
Magnetizing inductance.
Synchronous angular frequency.
Rotational angular frequency.
Slip frequency.
Number of poles.
Stator-flux vector.
Rotor flux vector.
Stator resistance.
Rotor resistance.
Slip.
Electrical torque.
Stator voltage vector.

Manuscript received June 19, 2012; revised October 5, 2012; accepted


December 20, 2012. Date of publication January 30, 2013; date of current
version February 28, 2013. This work was supported in part by the Fondo
Nacional de Ciencia y Tecnologa, Chile, under Contract 1110984 and Contract
1121104 and in part by the Industrial Electronics and Mechatronics Millennium
Nucleus.
R. Crdenas is with the Department of Electrical Engineering, University of
Chile, Santiago 837-0451, Chile (e-mail: rcd@ieee.org).
R. Pea is with the Department of Electrical Engineering, University of
Concepcin, Concepcin 407-4580, Chile (e-mail: rupena@udec.cl).
S. Alepuz is with the Matar School of Technology, Polytechnic University
of Barcelona, 08302 Matar, Spain (e-mail: alepuz@tecnocampus.cat).
G. Asher is with the Department of Electrical and Electronic Engineering,
University of Nottingham, Nottingham, NG7 2RD, U.K. (e-mail: greg.asher@
nottingham.ac.uk).
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/TIE.2013.2243372

HE DOUBLY fed induction machine (DFIM), also known


as the wound-rotor or slip-ring induction machine, is an
induction machine with both stator and rotor windings [1], [2].
The DFIM is nowadays widely used as a generator, particularly
in variable-speed wind energy applications with a static converter connected between the stator and rotor. Currently, this
topology occupies close to 50% of the wind energy market [3].
Table I shows some of the commercially available wind energy
conversion systems (WECSs), with power in the range of
1.53 MW, which are based on doubly fed induction generators
(DFIGs). In total, in Table I, there are 93 models of WECSs
based on DFIGs for that power range. In Table I, NM stands
for number of models.
DFIGs are also used in higher power ranges (> 3 MW).
The German company Repower manufactures two models of
WECSs based on DFIGs, the model 6M with a total output
power of 6150 kW and the model 5M with a total output power
of 5 MW [4].
For WECSs based on DFIGs, gearboxes are required because
a multipole low-speed DFIG is not technically feasible [5]. The
design of a DFIG-based WECS with a one-stage gearbox was
proposed in [6], but no commercial WECS has been implemented with this concept. However, even with the problems
associated with a three stage (3S) gearbox, the DFIG still has
some advantages when compared with other generators used in
wind energy applications [3]. For instance, in [7] and [8], three
generators suitable for wind energy applications are studied:
a direct-drive synchronous generator (SG) (which is one of
the solution offered by Enercon [9]), a direct-drive permanentmagnet generator (PMG) [10][14] (marketed by several companies, e.g., Vestas [13], Clipper [14], and Dewind), and a
3S-geared DFIG (see Table I). The results in terms of weight,
cost, size, and losses obtained in [7] and [8] are presented in
Table II. Notice that the 3S-Geared DFIG is considered the base
for the comparison.

0278-0046/$31.00 2013 IEEE

CRDENAS et al.: CONTROL SYSTEMS FOR OPERATION OF DFIGs IN WIND ENERGY APPLICATIONS

TABLE I
C OMMERCIALLY AVAILABLE WECSs IN THE
R ANGE OF 1.53 MW BASED ON DFIGs

2777

One of the main reason for the popularity of DFIGs in wind


energy applications is that relatively small power converters
are required to control the generator. For a typical DFIG, the
power converters are connected in the rotor circuit and, for
restricted speed range, are rated at a fraction (usually 30%) of
the machine nominal power [15][17]. Typically slip rings are
required in order to connect the machine-side converter to the
rotor. Brushless topologies are also feasible [18][22].
Because of the popularity of DFIGs for wind energy generation, control systems suitable for this application have been
extensively investigated. Control methods for grid-connected
WECSs, stand-alone systems, frequency support using DFIGs,
low-voltage ride-through (LVRT) control, etc., have been presented and discussed in the literature. The aim of this paper is to
give an update of the most recent trends regarding DFIG control
systems. In this respect, it augments previous overview papers
[23], [24]. In particular, this paper highlights the most recent
issues in sensorless control of DFIGs, droop control, the application of DFIGs to microgrids, and the latest work in LVRT.
This paper is organized as follows. In Section II, speed and
torque control of DFIGs is discussed, and the maximum power
point tracking (MPPT) control of DFIGs is also analyzed. In
Section III, control of DFIGs connected to unbalanced grids
is presented. Section IV addresses DFIG sensorless control
methods, whereas Section V discusses frequency support using
DFIGs. LVRT control is discussed in Section VI. Finally, the
conclusions are presented at the end of this paper.
II. S PEED AND T ORQUE C ONTROL OF DFIM
A. OptiSlip of Vestas

TABLE II
C OMPARISON B ETWEEN T HREE G ENERATORS P UBLISHED IN [7] AND [8]

From Table II, it is concluded that the total weight of a WECS


based on a direct-drive PMG is about 4.5 times higher than that
of a WECS based on a DFIG [7], [8]. The stator diameter of a
direct-drive PMG is about six times that of a DFIG of similar
power. Recently, the performance of the DFIG has been also
compared with that of the medium-speed permanent-magnet
SG (PMSG) in [8]. The medium-speed PMSG, usually coupled
to a single-stage gearbox, is a relatively new topology for
variable-speed wind generation (also known as the Multibrid
concept [3]) and has been adopted by some WECS manufacturers, e.g., Vestas, Areva, and WinWinD [3].

In the past, external resistors were connected to the slip rings


of wound-rotor machines in order to reduce the starting current
(for motor operation) or for maximizing the electrical torque in
a given operating point. The use of external resistors, at least for
these applications, is now considered obsolete because better
performance is obtained using power electronics as soft starters,
pulsewidth-modulated (PWM) inverters, etc. However, external
resistors connected to the rotor are still used in some topologies
of WECSs based on wound-rotor induction machines.
Vestas in its OptiSlip scheme [25][27] (e.g., Vestas V39
600, V661.65 MW) places the resistors and electronic components (as current sensors, insulated-gate bipolar transistors
(IGBTs), and part of the control hardware) mounted in the rotor,
i.e., no slip rings are required. Depending on the operating point
of the WECS, different ohmic values of resistors are connected
to the rotor windings using the IGBT transistors. The signals for
the control of the IGBTs are transmitted via an optical link from
outside the rotor. This topology is designed for a slip variation
of up to 10%, delivering a smoother power to the grid [25],
[26], [28]. Moreover, the mechanical stresses on some parts of
the wind turbines are drastically reduced [25].
A further development of the OptiSlip is the OptiSpeed
scheme, which allows slip variations of about 60% [28]. The
use of external resistors connected to the rotor could be augmented with pitch control in order to improve the performance
of the WECS in dynamic operation, e.g., in the presence of a
grid disturbance [27].

2778

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 7, JULY 2013

devices in the MCs could be a problem to obtain a good


performance for LVRT conditions. Further investigation about
this issue is required before considering the MCs as suitable
candidates for controllings DFIGs in WECSs.
C. Vector Control of DFIMs

Fig. 1. Static Scherbius scheme with two back-to-back PWM VSIs.

Even when the OptiSlip of Vestas has an important share


of the total WECSs installed in the world (11% in 2008
according to [27]), the main disadvantage of this topology is in
its relatively low efficiency because of the dissipation of energy
in the external resistors.
B. Static Scherbius Drive
The Scherbius system was proposed by the German engineer
Arthur Scherbius in the early years of the 20th century. The
scheme allows bidirectional power flow in the rotor circuit so
that operation of the machine below and above synchronous
speed is possible. Several topologies have been used in this
scheme [1], [2], [15], [17], [29][51]. The first work reported
in the literature uses a topology similar to the static Kramer
drive discussed in [52], but with the rotor diode bridge replaced
by a current-fed (naturally commutated) dc-link converter [32],
[44], [46], [53]. Another early topology of the Scherbius drive
uses a cycloconverter connected between the stator and the
rotor [43], [50], [54], However, the current-fed converters and
the cycloconverters produce high harmonic content in the rotor
current, which are reflected in the stator due to the transformer
action of the machine.
The disadvantages of the naturally commutated converters
can be overcome by the use of two PWM voltage-fed currentregulated inverters connected back to back in the rotor circuit
[15], [17], [29][31], [33], [35][42], [45], [47][49], [51],
[55]. The scheme is shown in Fig. 1. This topology allows:
bidirectional power flow with operation below and above
synchronous speed with the speed range restricted only by
the rotor voltage rating of the DFIG;
operation at synchronous speed with dc injected into the
rotor, with the rotor-side inverter operating in chopping
mode;
low distortion stator, rotor, and supply currents;
independent control of the torque and rotor excitation;
control of the displacement factor between the voltage and
the current in the GSC and, hence, control over the system
power factor.
The application of direct frequency power converters, namely
matrix converters (MCs), and indirect MCs (IMCs) has been
proposed as an alternative to the back-to-back voltage source
inverter (VSI) topology shown in Fig. 1 [56], [57]. These
topologies are all-silicon solutions for acac conversion with
sinusoidal input and output currents without using passive
components in the dc link. However, the lack of energy storage

The vector control technique developed for squirrel-cage


induction machines [1], [58] can be extended to DFIMs [15].
Usually, in a cage induction machine fed by an inverter connected to the stator, the stator currents are controlled using a
dq rotating frame aligned with the rotor flux. By analogy,
in DFIMs, the rotor is fed by an inverter; therefore, the rotor
currents are usually controlled using a rotating frame aligned
with the stator flux [15], [16]. Under this scheme, the electrical
torque is proportional to the q-axis rotor current. Because the
stator is connected to the utility in grid-connected applications,
the d-axis rotor current can be used to regulate the reactive
power flow in the machine. In wind energy applications, MPPT
is usually carried out by controlling the machine electrical
torque [15], [41], [55]. This is discussed in Section II-G.
The machine equations for a DFIG in a dq synchronous
frame orientated along the stator flux are as follows [1], [15]:

ds
qs
=

dr
qr


vds
=
vqs

ids
Ls 0 L0 0
0 Ls 0 L0 iqs

L0 0 Lr 0
idr
0 L0 0 Lr
iqr





d ds
Rs 0
ids
+
iqs
0 Rs
dt qs



0 e
ds
+
e
0
qs






d dr
Rr 0
idr
vdr
=
+
0 Rr
vqr
iqr
dt qr



0 sl
dr
+
sl
0
qr
p
Te = ktransf L0 (iqs idr ids iqr )
2

(1)

(2)

(3)
(4)

where subscripts d and q denote direct and quadrature components referred to the synchronous rotating frame, respectively;
and subscripts r and s denote stator or rotor quantities, respectively. s = L0 ims is the stator flux, where ims is known as
the magnetizing current.
The field orientation for machine variable transformation
uses slip angle sl derived from the position of the statorflux vector s and the rotor position r (see [15] and [16]) as
follows:
sl = s r .

(5)

The position of the stator-flux vector s can be obtained from


the stator-flux components as


s
s = tan1
.
(6)
s

CRDENAS et al.: CONTROL SYSTEMS FOR OPERATION OF DFIGs IN WIND ENERGY APPLICATIONS

2779

Fig. 3. Voltage vectors and stator and rotor fluxes.

Fig. 2.

dynamic performance of the controlled currents is similar, but


in this case, the d-axis and q-axis stator current components
are proportional to the stator active and reactive power [30].
Standard modulation techniques could be used to provide the
PWM patterns to the rotor-side converter (RSC)/GSC [63]
[65]. For parallel connection of power converters, shifting of
the PWM patterns could be implemented in order to reduce the
total harmonic distortion [64].

Control schematic of a DFIG.

An alternative to (6) is to use a phase-locked loop (PLL) [59]


to obtain s . The components of the stator flux can be
calculated from the stator voltages and currents as

s = (vs Rs is ) dt

(7)
s = (vs Rs is ) dt.
The expression in (7) requires an integrator. However, in
practical implementations, a pure integrator can be replaced by
a low-pass filter or a bandpass filter (BPF) used as a modified
integrator to block the dc component of the measured voltages
and currents [60], [61]. The BPF is typically designed with a
cutoff frequency of 0.1 to 1 Hz. Because the stator voltages
and currents are 50-Hz signals, the performance deterioration
from integral action is negligible [61]. The control schematic
is shown in Fig. 2, where E is the converter dc-link voltage,
and the superscript denotes a demand value. When the
orientation along the stator flux is correct, the electrical torque
is given by
Te = ktransf

p L20
ims Iqr = kt1 ims iqr
2 Ls

(8)

with the torque constant kt1 = ktransf pL20 /Ls . The stator magnetizing current ims = ds /L0 is practically constant in gridconnected applications. Under flux orientation conditions, the
magnetizing current can be provided: 1) entirely from the stator
with ird = 0; 2) entirely from the rotor with isd = 0; or 3) a
combination of magnetizing currents supplied from both the
stator and the rotor. This degree of freedom regarding the
reactive power flow in the machine can lead to an optimization
problem, where losses in the machine and ratings of the rotorside and the line-side converters need to be considered [41],
[62]. The electrical torque is proportional to irq , and the reactive
power in the machine can be regulated by acting upon ird .
Vector control schemes can also be implemented using a
reference frame oriented along the stator voltage vector and
controlling the stator currents instead of the rotor currents. The

D. DTC of DFIM
The direct torque control (DTC) technique [66], widely
applied to squirrel-cage induction machines, has also been used
to control the electrical torque in the DFIM because of the good
dynamic performance that it achieved [31], [39], [40], [45],
[67][69]. ABB has developed a low-voltage power converter
to control a DFIM for wind power applications using this
technique [70].
A two-level voltage-fed inverter can impose six active vectors and two zero vectors at the machine rotor terminals, as
shown in Fig. 3(a). These voltage vectors, when applied for
time interval t, produce changes in the rotor flux vector both
in magnitude and phase with respect to the stator-flux vector
(see also Fig. 3). It can be shown that the electrical torque is
proportional to the cross product of the stator and rotor flux
vectors, i.e.,
Te = kt s r = kt |s ||r | sin

(9)

where kt is a constant dependent on the machine parameters.


Assuming grid-connected operation, the stator-flux magnitude is practically constant. Therefore, the rotor flux vector
can be changed by applying different rotor voltages via the
RSC. From (9), this produces changes in the electrical torque
(and generated reactive power). Depending on the position of
the rotor and for a desired change in the electrical torque and
rotor flux magnitude, there is an optimum voltage vector to be
applied to the machine [2], [45].
In order to implement the DTC strategy, it is necessary to
know the rotor flux vector in magnitude and angle, and the
electrical torque. The rotor flux can be obtained using (1) with
the stator and rotor currents referred to the reference frame
affixed to the rotor, i.e.,

2 + 2
|r | = r
r
r2 = L0 irs + Lr ir ; r = L0 irs + Lr ir
r = tan1 (r /r )

(10)

2780

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 7, JULY 2013

Fig. 4. DTC for DFIM.

where irs and irs are the components of the stator


current referred to the rotor frame. The electrical torque can be
obtained as
Te = kt2 (r ir r ir )

(11)

where kt2 is dependent on the transformation being used.


The control diagram of a standard DTC strategy is shown
in Fig. 4.
E. DPC Applied to DFIMs
The direct power control (DPC) technique was proposed
about 15 years ago for controlling three-phase PWM rectifiers
[71][73]. It follows the same philosophy of DTC, but it also
looks at the effect of the stator and rotor fluxes upon the stator
active and reactive power. It can be shown that stator active
power is proportional to the rotor flux component perpendicular
to the stator flux where the stator reactive power is proportional
to the rotor flux component aligned with the stator flux [44].
The approach can be extended to DFIMs [29], [42], [45],
[49], [51], [74]. A DPC strategy minimizing the use of zero
voltage vectors is presented in [49]. When using DTC at low
rotational speed, zero voltage vectors are more frequently applied to the machine terminals causing a flux reduction because
of the stator resistance. In DFIMs, the equivalent situation is the
operation near or at synchronous speed where the rotor voltage
applied to the machine is low. Operation at or near synchronous
speed is not uncommon when the machine is used in variablespeed WECSs.
The operation principle is to control directly the stator active
and reactive power by applying the proper voltage vector in the
machine rotor. The stator power can be calculated as
Pe = ktransf (vs is + vs is )
Qe = ktransf (vs is + vs is ).

(12)

The error between the reference active and reactive power


and the calculated active and reactive power in the machine
are processed by hysteresis controllers. Schemes employing
both two-level and three-level hysteresis controllers have been
reported in the literature [45], [49]. The implementation of the
strategy needs the rotor flux vector position within six predefined sectors in the rotor coordinates in order to determine the
optimal rotor voltage vector to apply to the machine. Because
the rotor flux vector position needs to be known, the standard

Fig. 5.

DPC scheme as reported in [49].

DPC approach requires, as the standard DTC strategy, stator


and rotor current measurements. However, it is claimed in [71]
[73] that DPC is less dependent of the machine parameters.
In order to reduce the strategy parameter dependence, alternative schemes have been used. A strategy based on the statorflux position, which is referred to the rotor, and the effect of
the different voltage vectors upon the stator active and reactive
power is presented in [44]; therefore only stator voltage and
currents are measured. A schematic of this strategy reported in
[49] is shown in Fig. 5.
Another strategy to estimate the rotor flux position is presented in [40], where an adaptive mechanism based on the effect
of voltage vectors upon the reactive power variation is presented. However, the strategy requires a rather high sampling
frequency. Again, only the stator current and voltages need to
be measured in order to implement the DPC.
The control schemes presented in Section II-DF have been
very well documented in the literature. The strategies provide good overall performance, but it is not straightforward
to establish the superiority of one over the others. A fair
comparison would have to include dynamic-state and steadystate performances, current ripple content, and losses in the
converters. The vector control approach is based on the machine
model and is more parameter dependent: the implementation
complexity might be higher; currents, voltages, and position
need to be measured (although implementation without encoder
is feasible); and the current control dynamics are reasonable
with no high sampling frequency. On the other hand, DTC
implementation is simpler, even if it is a model-based approach,
and is less dependent on machine parameters: high torque
dynamics can be achieved, but higher nonconstant switching
frequencies are typical; a higher current ripple is expected,
higher bandwidth of current and voltage sensors are needed
and rotor position needs also to be measured. Finally, DPC
could be even simpler to implement: good power dynamics can
be achieved with high variable switching frequency; a higher
current ripple is usual and higher bandwidth of current and
voltage sensors are also needed, but rotor position does not need
to be measured.
If MPPT is considered, speed measurement/estimation is
typically required for any of the control strategies discussed
earlier. The MPPT implementation is straightforward for DTC
and vector control approaches because the electrical torque

CRDENAS et al.: CONTROL SYSTEMS FOR OPERATION OF DFIGs IN WIND ENERGY APPLICATIONS

2781

Fig. 6. Locus for the maximum aerodynamic efficiency, in the power-speed


plane, for a typical variable-speed wind turbine.

is directly or indirectly controlled. However, DPC controls


only the stator power, and MPPT requires regulation of the
total power supplied to the grid or isolated load. Therefore,
the rotor power has to be considered, and this increases the
implementation complexity of the MPPT algorithm.
Fig. 7. Some simple control system for MPPT in DFIG-based WECSs.
(a) Control system of (13) and (14), as discussed in [77] and [78]. (b) Control
system of (15), as discussed in [55] and [84].

F. Control of the GSC


The objective of the line-side converter or the GSC in the
topology depicted in Fig. 1 is to permit the active power flow,
regulating the dc-link voltage to a constant level. Close-tounity power factor operation is usual, but it is also possible to
control the reactive power flow between the converter and the
stator/grid. A vector control approach is normally used [15],
[55], with a reference frame oriented along the grid-voltage
vector, enabling independent control of the active and reactive
power flowing between the grid and the GSC. The grid-side
PWM converter is current regulated, with the d-axis current
regulating the dc-link voltage and the q-axis current regulating
the reactive power. Alternatively, DPC can be also applied to the
control of the GSC, leading also to a decoupled control of the
active and reactive power flows in the converter [29], [71][73].
G. MPPT Control
For a typical variable-speed wind turbine, the locus of the
maximum aerodynamic efficiency corresponds to a cubic line
relating the power captured with the rotational speed [75][81].
This is shown in Fig. 6. The optimal power Popt is related to
the rotational speed of the blades by the following nonlinear
function:
Popt = kopt r3

(13)

where kopt is a function of the parameters of the WECS, e.g.,


gearbox size, blade radius, blade profile, etc. Two types of
MPPT algorithms have been reported in the literature, i.e., the
speed control and torque control of the electrical generator for
maximum aerodynamic efficiency [82], [83]. For the MPPT
algorithms based on speed control, the generator rotational
speed is regulated to drive the WECS to the point of maximum
aerodynamic efficiency. Further discussion of MPPT methods
based on speed control is considered outside the scope of this
paper and the interested reader is referred elsewhere [82], [83].

For the MPPT algorithms based on torque control, the


quadrature current iqr is regulated to drive the WECS to the
point of optimal power capture. As discussed in [77] and [78]
to drive the WECS to the point of maximum aerodynamic
efficiency, the electrical torque could be controlled as
Te = kopt r2 .

(14)

Using (8) and (14), the quadrature reference current iqr can
be calculated as
iqr =

kopt 2
.
kt1 ims r

(15)

If the machine parameters are correctly identified, the simple


control strategy of (15) can be used to drive the WECS to the
point of maximum aerodynamic efficiency.
The control system based on (15) is shown in Fig. 7(a). The
rotational speed of the generator is used as the input of a lookup
table (or nonlinear function), where (15) is stored. Current iqr is
obtained at the lookup table output and is used as the reference
of the quadrature current control loop.
Another alternative is to implement optimal power tracking
using an additional control loop. This strategy has been reported
in [55] and [84]. The control system calculates the power
reference Pe using a lookup table, where the optimal power as
a function of the rotational speed is stored [see Fig. 7(b)]. From
Pe , the rotor torque current is calculated as

iqr = Kp (Pe Pe ) + ki (Pe Pe ) dt


(16)
where kp and ki are the proportional and integral constants
of the proportionalintegral (PI) controller. Pe is the electrical
power supplied by the DFIG to the grid, which is measured by
the voltage and current transducers. The control system shown
in Fig. 7(b) requires nested control loops with the bandwidth of
the outer loop being a fraction of the internal current iqr loop.

2782

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 7, JULY 2013

Fig. 8. DFIG feeding a stand-alone load.

The main advantage of the control strategy of (16) [and


Fig. 7(b)] is that the errors in the machine parameter estimation,
e.g., kt1 and ims in (15), are compensated by the PI controller.
On the other hand, the relationship between the torque rotor
current and power Pe is dependent on the rotational speed,
and some compensation strategy, for instance gain scheduling
control, could be required to maintain a good dynamic response
in the whole operating range.
Here, only two simple control strategies have been explained. However, other power tracking methodologies (as the
speed-control-based MPPT algorithms discussed in [77] and
[85][87]) e.g., perturbation and observation, wind speed observers, etc., can be applied to WECSs based on DFIGs.
III. C ONTROL S YSTEMS FOR THE C ONNECTION
OF DFIG S TO U NBALANCED S YSTEMS
A WECS may be installed in remote rural areas, where weak
grids with unbalanced voltages are not uncommon [42], [88]
[90]. Moreover, in stand-alone applications, the DFIGs can feed
unbalanced and islanded loads [91][95].
As reported in [89], [90], and [96][99], induction machines
are particularly sensitive to unbalanced operation since localized heating can occur in the stator, and the lifetime of
the machine can be severely affected. Furthermore, negativesequence currents in the machine produce pulsations in the
electrical torque, increasing the acoustic noise and reducing the
life span of the gearbox, blade assembly, and other components
of a typical WECS [88], [91], [92], [100].
For the control of DFIGs operating in unbalanced systems,
control algorithms based on counterrotating synchronous dq
axes [89], [91], [92], [98], [101], resonant control [95], [102],
[103], predictive control [93], [104], [105], sliding control
[106], and DPC [29], [42] have been proposed in the literature.
A. Control of a DFIG Feeding a Stand-Alone
Unbalanced Load
Fig. 8 shows a DFIG feeding a stand-alone load. The
DFIG stator and the load are star-connected with the neutral
points connected, to provide a path for the circulation of zerosequence currents. A four-leg grid-side inverter can be also
used to supply zero-sequence signals to a star-connected linear/
nonlinear unbalanced load [107][109].
The initial excitation for the system start up could be provided by a battery bank (not shown in the figure). The battery
could be kept charged afterward using the energy flow in
the dc link. Another possibility is to use a bank capacitor in
the stator for the self-excitation of the machine, generating the

Fig. 9.

Control system discussed in [91].

required stator voltage. Then, the control strategy of the lineside converter or, in this case, the stator-side converter, could
regulate the required dc-link voltage.
To compensate the load unbalance, the GSC and/or the RSC
can be used. For instance, in [91] and [92], the use of the GSC
to compensate the load unbalance is proposed. The control
system discussed in [91] is shown in Fig. 9. (Only the GSC
control system is shown.) The positive-sequence vector control
system is oriented along the stator voltage vector. Because of
the unbalance, a PLL is implemented to calculate the stator
voltage angle v [59]. From +v and v , the currents can
be referred to two synchronous d-axis and q-axis rotating at
+e and e , respectively. Doubly frequency components
are produced when the positive/negative-sequence currents are
referred to the d-axis and the q-axis rotating in the opposite
direction. As shown in Fig. 9, notch filters are used to eliminate
these high-frequency components [89], [91], [92], [98], [101].
+

The output of theses filters are the currents i+


df , iqf , idf , and iqf .
The control systems for the front-end positive-sequence cur+
rents i+
df and iqf are entirely conventional (see Fig. 9, and
[91] and [92]). Current i+
df regulates the dc-link voltage E, and
+
current iqf regulates the reactive power supplied to the load.
The front-end negative-sequence currents are regulated to



i
=
i
=

i
+
i
(17)
dqf
dqL
dqs
dqf .
Therefore, the negative-sequence current demand is a function
of the load negative-sequence current. In the steady state, when

i
dqf = idqL , stator current idqs = 0 [see (17)], and the torque
pulsations are eliminated.

CRDENAS et al.: CONTROL SYSTEMS FOR OPERATION OF DFIGs IN WIND ENERGY APPLICATIONS

2783

Fig. 11. DFIG feeding an unbalanced grid.

is controlled to eliminate the negative-sequence currents from


the stator of the DFIG.
B. Control of a DFIG Feeding an Unbalanced Grid
Fig. 10. Experimental results corresponding to the control system discussed
in Fig. 9. (a) Negative-sequence currents. (b) Stator and rotor unfiltered currents
referred to the dq positive-sequence axes.

In abc coordinates, the total voltage demand for the front-end


converter is obtained as (see Fig. 9)
+

+ vabc
.
vabc = vabc

(18)

Fig. 10 shows the performance of the control system depicted in Fig. 9 for negative-sequence current compensation
under variable-speed stand-alone operation (see [92]). The load
consists of three unbalanced resistors connected to phases a,
b, and c, respectively (see Fig. 8). The rotational speed is
varied from 1350 to 1650 rpm to illustrate the performance
at variable speed (from below to above synchronous speed).
Before t 1.25 s, the compensation system is not operating,
and the stator current has a negative-sequence component [see
Fig. 10(a)]. In t 1.25 s, the compensation is enabled, and the

stator current i
dqs is driven to zero. For t > 1.5 s, idqL idqf ,
and the negative-sequence currents are eliminated from the
machine stator. Notice that the term unfiltered indicates that
the displayed currents have not been filtered by the notch filters
shown in Fig. 9.
There are other publications where control of a stand-alone
unbalanced load is discussed. For instance, the control system
discussed in [94] uses the RSC to regulate a balanced load
voltage. In this study, the control system is tested with nonlinear
loads and the authors claim a good performance. However, the
main disadvantage of [94] is that the rotor current reference has
negative-sequence components, and a relatively large dc-link
voltage could be required to regulate these components.
Predictive control systems for DFIGs feeding unbalanced
stand-alone loads are discussed in [93] and [105]. In this case,
the voltage vector that minimizes a cost function is identified
and applied to the RSC. The control discussed in [93] and [105]
use only the RSC to compensate the unbalances in the standalone load.
To the best of our knowledge, the only publication reporting
the use of both the RSC and GSC to compensate the load
unbalance in a stand-alone DFIG is [95]. In this case, a dq control system augmented with a resonant controller (implemented
in the synchronous rotating frame) is proposed. The RSC is
controlled to regulate a balanced load voltage, whereas the GSC

Fig. 11 shows a DFIG feeding an unbalanced grid. In this


case, the aim of the control system is no longer to regulate
the grid voltage. The control approach shown here can be
useful to meet the LVRT requirements, as will be discussed
in Section VI. For unbalanced conditions, neglecting the zerosequence components in the system, the stator and current
voltage vectors can be written as [98], [103]
v s = v1s eje t + v2s eje t

(19)

is = i1s eje t+1s + i2s eje t+2s .

(20)

The voltage and current vectors at the output of the GSC (see
Fig. 11) can be written as
v f = v1f eje t+1f + v2f eje t+2f

(21)

if = i1f eje t+3f + i2f eje t+4f

(22)

where the subscripts 1 and 2 are used to indicate signals of


positive and negative sequences, respectively. Angles if and
is indicate a phase angle shift with respect to the stator voltage
angle. The active and reactive power supplied by the DFIG to
the grid can be calculated from [29], [42], [103]


(23)
S = ktransf v f icf + v s ics .
In (23), the superscript c is used to indicate the complex
conjugate operator.
It is relatively simple to show that the active power and
reactive power of (23) have three terms: the mean value, a term
proportional to sin(2e t), and a term proportional to cos(2e t)
[98], [101], [102], [110]. This can be written as
P = Pavg + Psin(2e t) + Pcos(2e t)

(24)

Q = Qavg + Qsin(2e t) + Qcos(2e t).

(25)

Using (19)(25), several control targets for the operation of


DFIGs in unbalanced grids can be defined [29], [42], [98],
[101], [103], [110]. For instance, in [103], one of the following
control targets is proposed:
To eliminate the oscillations in the total active power output from the overall system, i.e., Psin(2e t) + Pcos(2e t)
in (24);

2784

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 7, JULY 2013

To reduce the oscillation in the total reactive power


supplied to the network, i.e., Qsin(2e t) + Qcos(2e t) = 0
in (25).
To supply a grid current with no negative-sequence
component, i.e., i2s eje t+2s + i2f eje t+4f = 0 [see
(20) and (22)].
Each power converter has four degrees of freedom allowing
the independent regulation of the (or dq) components
of the negative and positive-sequence output currents. In some
papers related to the control of DFIGs connected to unbalanced
systems, only one of the converters is used. For instance, in
[6], the GSC is used to compensate the negative-sequence
current of the load. On the other hand, in [29], [42], [88], [89],
[94], and [99], only the RSC is used to compensate the grid
unbalance. DPC [29], [42] and dq control are proposed in
these publications to compensate the grid unbalance, injecting
negative-sequence currents in the rotor.
In recent papers, the control of both the GSC and the RSC
has been proposed to compensate the grid unbalance. This has
the advantage that additional degrees of freedom are introduced
in the control system by using two power converters, and more
control targets can be achieved [98], [101][103], [111], [112].
IV. S ENSORLESS C ONTROL OF DFIGs
The DFIG can be used as a variable-speed generator in standalone and grid-connected applications [95], [113][116]. In
both cases, the use of sensorless vector control is desirable
because position encoders or speed transducers have many
drawbacks in terms of maintenance, cost, robustness, and cabling between the speed sensor and controller [78].
There are several sensorless methods reported in the literature. In this paper, they are classified as open-loop sensorless
methods, model reference adaptive system (MRAS) observers,
and other sensorless methods. Most of the sensorless control
methods reported here have been applied to conventional vector control of DFIGs (see Section II-C). However, sensorless
schemes can also be applied to the control methods discussed
in Section II-D and E.

Fig. 12.

Block diagram of a typical MRAS observer.

In (26) and (27), irs is the rotor current vector referred to the
stator, and (ir + jir ) is the measured rotor current in
coordinates. Using (26) and (27), an estimation of the slip angle
is obtained as


(28)
sl = tan1 (ir /ir ) tan1 isr /isr .
Using (28) in (5), an estimation of rotor position angle is
derived. Open-loop methods are not only based on estimation
of the DFIG rotor current vector. In [122], an observer based
on the magnetizing current derived from the rotor and stator
equations of the machine is proposed, although only simulation
results were presented, and no methodology was proposed
for the observer modeling and design. In [123], a rotor-fluxbased sensorless scheme is proposed, where the rotor flux
is obtained by integrating the rotor back electromotive force.
This sensorless method has poor performance when the machine is operating around the synchronous speed because
the rotor is excited with low frequency voltages. Therefore,
the rotor flux cannot be accurately estimated by integrating the
rotor voltages.
In the open-loop methods, the rotational speed is obtained
via differentiation of the estimated slip angle of (28), which can
amplify the high frequency noise. Moreover, for the open-loop
methods reported in the literature, issues of observer modeling,
observer bandwidth, and design methodology for the whole
sensorless system are not discussed.
B. Sensorless Method Based on MRASs

A. Open-Loop Sensorless Methods


Most of the early work in sensorless control of DFIGs is
based on open-loop methods, where the estimated and measured rotor currents are compared in order to derive the rotor
position [117][121]. For instance, the rotor current referred to
the stator can be estimated using the stator flux and the stator
current as [121]
sr =

( s Ls s )
L0

(26)

The measured rotor current can be referred to the stator using


isr = (ir + jir )ejsl
where the slip angle is defined in (5).

(27)

The MRAS was first introduced for sensorless control of cage


induction machines in [124]. In this publication, the observer
design is discussed, and a small-signal model is proposed. Most
of the MRAS observers proposed in the literature for cage
induction machines are based on rotor flux estimation.
The application of MRAS observers for sensorless control of
DFIGs was first reported in [125] and [126]. However, in these
papers, only simulations were presented for a DFIM operating
at very low rotational speed. Issues such as observer dynamics,
control design procedure, sensorless accuracy, and sensitivity
to machine parameter variations were not addressed. Further
publications discussing the application of MRAS observers for
sensorless control of DFIGs were presented in [127] and [128].
In the general case, an MRAS observer is based on two
models [61], [77], [124], [129][134]: a reference model and
an adaptive model (see Fig. 12). The estimated speed and rotor
position are used to adjust the adaptive model, driving the error

CRDENAS et al.: CONTROL SYSTEMS FOR OPERATION OF DFIGs IN WIND ENERGY APPLICATIONS

2785

Fig. 14. Sensorless control of a grid-connected DFIG using a stator-fluxbased MRAS observer. Notice that the control system is unstable when the
rotor magnetizing current idr0 is driven to zero.
Fig. 13. (a) Stator-flux-based MRAS observer proposed in [127]. (b) Smallsignal model corresponding to the stator-flux MRAS observer.

to zero. This error is usually defined as the cross product


.
between reference vector x and derived estimated vector x
Mathematically, this can be written as
=x
d xq xd x
q = |x||
x| sin( )

(29)

where is the phase angle between the vectors (


x, x). In
[115], the small-signal model, machine parameter sensitivity,
and the design procedure of a stator-flux MRAS observer were
presented (i.e., x = s ). In this case, the reference model and
the adaptive model are obtained, respectively, as

(30)
s = (v s Rs is )dt

s = Ls is + L0 ir ej r .

(31)

The estimated rotor position angle r is used to drive the


error of (29) to zero. The implementation of the stator-flux
MRAS observer is shown in Fig. 13(a), and the small-signal
model corresponding to this observer is shown in Fig. 13(b).
As shown, the gain of the feedforward path is dependent on the
magnetizing current idr0 . Therefore, if the DFIG is operating
in a grid-connected application and if the magnetizing current
required for the generator is entirely supplied from the grid,
then the rotational speed cannot be tracked by the observer.
The experimental result depicted in Fig. 14 further corroborate the small-signal model in Fig. 13(b) [133]. The DFIG is a
sensorless vector controlled using a stator-flux MRAS observer;
when t = 23 s, the magnetizing rotor current ird0 is driven to
zero and the system becomes unstable because tracking of the
rotor position angle and rotational speed is lost.
In [61], [136] a rotor current MRAS observer (RCMO) is
proposed, which is appropriate for grid-connected and standalone operation for most of the DFIG operating range. In
this case, the reference model is simply the measured rotor
current. The adaptive model is derived from (26) and can be
written as
r =

( s Ls s )
L0

ej sl .

(32)

Fig. 15. (a) RCMO presented in [61]. (b) Small-signal of the RCMO presented in [61].

A detailed description of the RCMO, including the methodology required to synchronize the DFIG to the grid, the smallsignal model, and the control algorithm used for catching the
speed on the fly, is presented in [61]. The implementation of
the RCMO is shown in Fig. 15(a). Fig. 15(b) shows a linearized
model of an RCMO, which is used to design the PI controller
in Fig. 15(a).
Experimental results obtained with a DFIG vector controlled
using a sensorless scheme based on an RCMO are shown in
Fig. 16. Fig. 16(a) shows the performance of the control system
used to synchronize the DFIG to the electrical grid before the
grid-connected generation is started. Notice that, in t = 20 s,
the power switch is closed, and the DFIG stator is connected
to the grid. Fig. 16(b) shows the experimental results obtained
for speed catching on the fly with sensorless control using the
RCMO. These experimental results are fully discussed in [61]
and [136].
From the small-signal model in Fig. 15(b), it is concluded
that the gain of the feedforward path is only affected by the
magnitude of the rotor current vector, which is not zero in the
typical operation range. Therefore, unlike the stator-flux MRAS
observer, the RCMO can be applied to sensorless control of
DFIG when the machine is grid connected and entirely magnetized from the stator. In fact, the RCMO can be applied to standalone and grid-connected application. In addition, as presented
in Fig. 16(a), sensorless vector control of the DFIG using an
RCMO is appropriate to synchronize the DFIG to the grid.

2786

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 7, JULY 2013

in [139] show that the variation in the stator inductance can be


compensated when the proposed adaptive algorithm is properly
designed.
A new sensorless control topology, which is also based on
the MRAS observer, is presented in [131] and [135]. The
proposed observer is called the torque-based MRAS observer
(TBMO) and uses a different methodology for estimating the
rotor current vector. Assuming that the vector control system
is orientated along the stator flux, then the torque and flux
components of the rotor current vector can be calculated as
irq =

s Te
s | is |
L
L
s
=
0 | |
0 | |
L
L
s
s



ird =
Fig. 16. Experimental results discussed in [61] and [136] corresponding to the
operation of a RCMO. (a) Synchronization to the grid. (b) Speed catching on
the fly.

Considering that the RCMO is able to operate in most of the


conditions required, i.e., grid-connected operation, stand-alone
operation, etc. [133], it is considered that the MRAS observer
has the best overall performance among the three sensorless
topologies discussed in that publication.
A variation of the RCMO is presented in [132]. Because the
error of (29) is a nonlinear function, in [132], it is proposed to
calculate the error using
= tan1 ([r ir ]/[r .ir ])

(33)

where r ir represents the cross product between the rotor


current and that estimated using a Luenbenger observer [132],
[137], [138]. On the other hand, r ir represents the inner
product between both currents. The use of (33) as normalized
error, instead of (29), has the advantage of producing a linear
model, where the error is proportional to the phase shift angle
between () and (ir ), instead of being proportional to the
nonlinear function sin( ) [see (29)]. This simplifies the design
of the PI controller in Figs. 13 and 15, enhancing dynamic
performance across the operating range.
The performance of the MRAS observers reported in the
literature depends strongly on the correct identification of the
inductances of the DFIM. In particular, the implementation of
the RCMO requires the correct identification of the magnetizing
and stator inductances [61], [133], [136], [139]. For gridconnected operation of DFIGs, the stator voltage can fluctuate about 10% of its nominal value, changing the level of
magnetic saturation in the machine. Therefore, the magnetizing,
stator and rotor inductances are subjected to variations. In order
to maintain the tracking of the rotor position angle, even in
the presence of grid-voltage fluctuations, adaptive tuning of the
stator inductance is proposed in [139]. This algorithm is based
on the fact that the magnitude values of the estimated rotor
current and that measured by the transducers are equal when the
machine parameters are correctly tuned, i.e., when |ir | = |r |;
s , and L0 = L
0 (||) is the rotor current vector estithen, Ls = L
mated from (32) (see [61]). The experimental results discussed

i2r + i2r i2rq

(34)

1/2
.

(35)

The dq components of the rotor currents calculated using


(34) and (35) are the reference model of the MRAS observer.
Note that the torque and flux rotor currents of (34) and (35) can
be calculated from the components of the stator flux and
measured stator/rotor currents.
According to [135], the main advantage of the proposed
TBMO is that (34) and (35) can be used directly as feedback
signals of the control system, increasing the dynamic response
and improving the stability of the whole system. Note that (34)
and (35) are not affected by errors in the rotor position angle.
C. Other Sensorless Methods for DFIMs
Sensorless control of DFIGs based on PLLs is proposed in
[141][143]. As discussed in [124], the operating principle of
PLLs is similar to MRAS observers because the error of (29)
and (33) are driven to zero when the phase shift between the
estimated vector and the reference vector is null. Therefore, the
sensorless observers of [141][143] have a similar performance
to those reported in [133]. However, some issues, such as the
design of the PI controller located in the PLL system and the
bandwidth of the rotor position observer, are not addressed in
[141] and [142].
Sensorless control of DFIMs is also discussed in the paper
reported in [144][147]. In [146] and [147], a rotor position
observer similar to the RCMO of [61] is discussed. The PI controller (see Figs. 1315) is replaced by a hysteresis controller.
The authors claim that this controller improves the performance
of the observer because the design of the hysteresis controller
does not require good knowledge of the plant parameters. However, as it is well known, controllers based on hysteresis may
produce signals of variable frequency at its output. Therefore,
some knowledge of the plant is usually required in order to
maintain this frequency inside a given operating range.
In [144] and [145], a rotor position observer, which is based
on the air-gap active and reactive power, is proposed. This
algorithm has some similarities to the TBMO reported in [135]
because the torque and flux components of the rotor current
are calculated using the components of the measured
voltage and currents without requiring an estimation of the

CRDENAS et al.: CONTROL SYSTEMS FOR OPERATION OF DFIGs IN WIND ENERGY APPLICATIONS

rotor position angle. Assuming a stator-flux orientation, the


dq components of the rotor current can be calculated as
dr =

Pg
Qg
=
|v s Rs is | qr
|v s Rs is |

(36)

where Pg and Qg are the power transferred across the air gap.
The current estimated using (36) are used in a RCMO in order
to estimate the rotor position angle. It is claimed in [144] and
[145] that the main advantage is that the calculation of the
stator-flux vector is not required in the vector control system.
However, in order to calculate Pg and Qg , an estimation of the
iron losses and magnetizing reactive power is required [144].
This can produce some errors, particularly when the DFIM is
operating with light loads.
Sensorless control of DFIMs can be also achieved using signal injection [148]. This methodology is relatively well known
for cage induction machines [149]. However, to the best of our
knowledge, sensorless control of DFIMs using signal injection
has only been discussed in [148]. The operating principle is
that the DFIM is a transformer in which the relative position
between the primary and secondary winding changes as the
rotor rotates. Therefore, if a high-frequency signal is injected
into the rotor, the phase of the corresponding signal in the stator
has a component that is dependent on the rotor position angle.
The main advantage of this method is high robustness against
variation in the machine parameters. However, experimental
validation of this method has not been reported, and injection
of high-frequency signals in the DFIG rotor is not simple to
achieve in relatively large machines, such as the ones used for
wind power generation.
To the best of our knowledge, the performance of the reviewed sensorless methods has not been studied for operation
in unbalanced grids or when the DFIG is feeding a stand-alone
unbalanced load of linear/nonlinear nature. Moreover, sensorless control of DFIGs during LVRT conditions has not been
addressed in the literature. Further research in these subjects is
required.
V. F REQUENCY S UPPORT U SING DFIGs
As wind power penetration increases, the fluctuating behavior of the wind velocity has more impact on the grid frequency.
Wind energy penetration may increase during periods of low
loads, e.g., in the night. In this case, grid-frequency fluctuations
above the maximum allowed by the grid codes can be produced
[150][152] if conventional MPPT is used to control the power
generated. In some countries, such as China [153], [154], about
27% of the yearly wind energy is curtailed because most wind
farms are operating using MPPT control without frequency
regulation [153]. In the past, the control used was based on
disconnecting part of the wind farm. Now, modern control
methods based on droop control and inertia emulation are
preferred [155].
Grid connection requirements (GCRs) are introducing regulations to establish grid-frequency support from wind turbines
[156]. For instance, according to the E-ON GCR [157], when
the frequency exceeds the value of 50.2 Hz, wind farms must
reduce their active power with a gradient of 40% of the available

2787

power per hertz, with a ramp rate of 10% of the grid connection
capacity per minute. A more detailed description about the
GCRs is found in Section VI-A.
There are several publications related to the subject of gridfrequency support using wind energy systems [153][155],
[158][170]. Most of the proposed methods use the kinetic
energy stored in the wind turbine rotating mass to provide additional power to the system in case of grid-frequency variation.
In power systems, inertia constant H is used instead of
inertia J. Constant H is defined as [155]
H=

Jr2
Ek
=
S
S

(37)

where S is the nominal apparent power of the WECS, and Ek


is the kinetic energy stored in the rotating blades. As shown in
(37), H is equal to the time that a WECS can supply the nominal
power using the kinetic energy stored in the rotor. The inertia
constants for WECSs are in the range of 26 s, whereas H for
a typical power system generator is in the range of 29 s [155].
Frequency support is usually accomplished using inertia
emulation and/or droop control. The output power of the DFIG
is controlled as a function of the grid frequency, i.e.,

= Pref + Kd (fgrid fref ) + Kei


Pout

d(fgrid )
dt

(38)

where Pref represents the output power demand for normal steady-state operation of the power system when the
grid frequency fgrid is equal to the reference frequency.
This power demand might be obtained, for instance, from a
lookup table, where a relationship between the rotational speed
and the demanded output power is stored. The second term
Kd (fgrid fref ) represents the droop power. In a typical
system, when the power is unbalanced, (e.g., there is more
or less consumption than power generation) the grid frequency changes. In this case, the DFIG output power is increased/decreased in order to support the generation. The last
term Kei (d(fgrid )/dt) corresponds to the inertia emulation. In
this case, the power demand is varied according to the rate
of change of the grid frequency. This component emulates
the inertia response of a conventional synchronous machine.

is regAssuming stator-flux regulation, reference power Pout


ulated using the quadrature rotor current in a DFIG, which is
controlled by the RSC.
To implement (38), the variable-speed WECS must have a
power reserve. Depending on the operating point, a combination of speed control and pitch control has been used to maintain
this reserve [153], [166][169]. In [79], the operating range is
divided into low-, medium-, and high-wind-speed sectors.
At low wind speed (e.g., 0 < V < V3 in Fig. 17), the
steady-state system is operating at a suboptimal power line,
for instance, at 90% of the maximum power curve shown in
Fig. 17. When the frequency decreases below fref , the generated power is increased by decreasing the rotational speed
until the maximum power point is reached (located in the curve
Pout = Kopt r3 ). If the grid frequency increases above fref , the
captured power is reduced by increasing the rotational speed.
At medium wind speed (e.g., V3 < V < V5 in Fig. 17), a
combination of speed control and pitch control is used. When

2788

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 7, JULY 2013

Fig. 17. Optimal and suboptimal power curve for the control strategy proposed in [79].

the turbine velocity reaches maximum speed, pitch control is


activated to avoid overspeeds.
At high wind speed, the power is regulated mainly by pitch
control. In this case, the output power is controlled below the
nominal value in order to maintain a power reserve, which is
used when the grid frequency goes below the reference value.
A control system for frequency support, also dividing the wind
speed into three operating areas, is presented in [167] and [169].
The main difference to that discussed in [79] is at low wind
speed; here, it is suggested to regulate the output power linearly
with the rotational speed, i.e.,

= kr .
Pout

(39)

According to [171], smoothed power generation could be


obtained when the rotational speed is linearly changed with the
power.
The application of variable-speed WECSs based on DFIGs
for frequency and voltage regulation in microgrids and minigrids has also been discussed [159], [172][174]. In this case,
the DFIG stator voltage is regulated according to
fs = kp (P P )

(40)

Vs = kq (Q Q)

(41)

where fs and Vs are the frequency and stator voltage of the


DFIG and kp and kq are the droops. A control system similar
to that proposed in [159] is shown in Fig. 18. The magnetizing
current supplied from the RSC is regulated to control the stator
voltage, and the stator frequency is varied according to (40).
An energy storage system (ESS) is used to supply power to the
grid or absorb excess power captured from the WECS. When
the ESS is fully charged, pitch control is required to limit the
power transferred to the grid.
VI. LVRT W ITH DFIGs
A. GCRs
In the last two decades, the installed wind power capacity has
considerably grown. At the end of 2011, the total installed wind
power world capacity reached 238.5 GW [175]. At the same

Fig. 18. Control system similar to that proposed in [159] for the operation of
DFIGs in microgrids.

time, wind energy penetration into the grid has significantly


increased. A good example is Spain, where the average wind
energy penetration has been 11%, 13.8%, and 16% in 2008,
2009, and 2010, respectively [176][178], although the wind
power penetration can temporarily reach a much higher value,
e.g., the 64% experienced on September 24, 2012 [179] in the
Spanish grid.
The GCRs are set by the power system operators to ensure
the reliability and efficiency of the utility [156], [180]. These
requirements can be divided into two main classes: steadystate or quasi-stationary operation requirements, and LVRT
requirements. A review of the GCRs of several countries is
presented in [156].
In steady-state or quasi-stationary operation, the requirements such as reactive and active power regulation to support
the utility voltage and frequency are specified in the GCR, and
have been dealt with in part in Section V.
Under grid disturbances, the former GCRs allowed the disconnection of the WECSs to avoid large overcurrents. However,
with the increase in the wind energy penetration, the sudden
disconnection of WECSs can lead to instability of the entire
power system [181], [182]. In this scenario, the power system
operators have updated their GCRs, and the wind generators are
required to remain connected to the grid during disturbances as
it is standard for conventional generators [156], [157], [180],
[183], [184].
With the current GCRs, the LVRT requirement demands
wind power plants to remain connected when a grid-voltage
sag occurs, thus contributing to maintaining stable network
voltage and frequency by delivering active and reactive power
to the grid with a specific profile depending on the grid-voltage
dip depth. Hence, LVRT is probably the most challenging
requirement among the GCRs, at least from the point of view
of the WECS.
LVRT requirements, extracted from the GCR of the utility
operator E-ON [157], are shown in Figs. 19 and 20. Very
similar curves are provided in the LVRT requirements of other
power systems operators [156], [180], [183], [184]. When a

CRDENAS et al.: CONTROL SYSTEMS FOR OPERATION OF DFIGs IN WIND ENERGY APPLICATIONS

2789

Fig. 21. Machine model from the rotor side.

Fig. 19. Voltage limit curve to allow generator disconnection.

Most of the disturbances are asymmetrical. Only 12% of grid


dips are symmetrical [189], [190].
As shown earlier, conventional regulation for the DFIG is
achieved by controlling the rotor currents. The machine model
seen from the rotor side is shown in Fig. 21 [182], [186],
[188], [191], where = (1 L20 /Ls Lr ). To control the rotor
currents by the RSC voltage, it is useful to calculate the opencircuit rotor voltage v rr0 . Note that superscript r denotes the
variables expressed in the rotor reference frame.
Considering the Park model for the induction generator
(1)(3) and the rotor in open circuit, the expression for the stator
flux is
d
Rs
= vs

dt s
Ls s

Fig. 20. Reactive current to be delivered to the grid under a voltage dip.

grid-voltage sag appears, the power generation plant must


remain connected to the grid if the line voltage remains over
the limit line 1 in Fig. 19 (region A). In certain cases, a brief
disconnection is allowed if the line voltage lies between the
limit-lines 1 and 2 (region B). Here, resynchronization typically
within 2 s is required to ensure a minimum reactive power
supply during the fault; also required is an active power increase
rate of > 10% of the rated generator power per second after
fault clearance [157]. A brief disconnection is always allowed
in region C, where resynchronization times of more than 2 s
and an active power increase following fault clearance of less
than 10% of the rated power per second are also possible in
exceptional cases. If the grid voltage remains low for longer
than 1.5 s (region D), selective disconnection of generators
depending on their condition can be carried out by the grid
protection system [157], [185]. In addition, during the voltage
sag, the WECS has to deliver a reactive current, specified in
Fig. 20, to aid the utility in holding the grid voltage. The
reactive power to be injected depends on grid-voltage reduction
during the dip, the system rated current, and the reactive current
given to the grid before the dip appears.
B. DFIG Behavior Under Grid Fault
A number of studies concerning the impact of grid faults
on DFIGs have been reported. For the grid, symmetrical disturbances, particularly the deep voltage sags, can be seen as
more stressing than asymmetrical disturbances since all phases
are lost. However, the analysis for asymmetrical disturbances is
more complex due to the appearance of negative-sequence components in the voltages and currents [3], [186][188]. DFIGs
have low negative-sequence impedance, and small negativesequence stator voltages can lead to high stator currents [106].

(42)

where the stator voltage can be expressed as the sum of the


positive (v1s ), negative (v2s ), and zero (v0 ) sequences
v s = v1s eje t + v2s eje t + v0 .

(43)

The solution for (42) is shown in (44). The zero-sequence


voltage does not create flux [188], [191]. From (44), the expression for the open-circuit rotor voltage shown in Fig. 21 can
be obtained, as shown in (45), which is given in the following:
s = 1s eje t + 2s eje t + n0 et/s
=

v1s je t v2s je t
e
+
e
+ n0 et/s
je
je
L0 jse t
L0
se
+ v2s (s 2)ej(2s)e t
Ls
Ls


L0 1
+
+ jr n0 et/sejr t .
Ls s

(44)

v rr0 = v1s

(45)

In normal operation, the grid voltage presents only the positive sequence, and the second and third terms in (44) and (45)
are zero. However, when a grid-voltage sag appears, the flux is
expressed as the sum of three components [182], [187], [188],
[191]: 1) the nonhomogenous or forced flux composed by two
terms corresponding to the positive- and negative-sequence
stator voltages; and 2) the homogenous or natural flux.
The natural flux vector does not rotate. This is a transient dc
component flux that exponentially decays with time constant
s = Ls /Rs and initial value n0 , which depends on the type
and depth of the grid-voltage sag and, in case of asymmetrical
dips, on the instant of time within the grid-voltage period in
which the grid disturbance occurs [182], [191].
The forced flux is the sum of the positive-sequence flux that
rotates at synchronous speed and the negative-sequence flux
[86]. The difference between asymmetrical and symmetrical

2790

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 7, JULY 2013

TABLE III
P OSITIVE , N EGATIVE , AND NATURAL F LUXES (P ER U NIT )
FOR D IFFERENT T YPES OF FAULTS [191]

voltage sags is the presence or absence of the negative-sequence


voltage and flux.
With respect to the rotor, as shown in (45), the open-circuit
rotor voltage has three components: 1) the positive-sequence
voltage rotating at se (the only component that is present in
balanced operation); 2) the negative sequence that rotates at
almost twice the synchronous speed (2 s)e and that only
appears when the disturbance is asymmetrical; and 3) the rotor
voltage produced by the natural flux that creates an open-circuit
voltage rotating at r .
When a grid disturbance occurs, the open-circuit rotor voltage has a large transient overvoltage (mainly caused by the
natural flux), which can be even greater than the stator voltage
[156], [182]. Because of the transient nature of the natural flux,
during symmetrical disturbances, the rotor voltages return to
positive-sequence values even if the grid fault is permanent [see
(45)]. During asymmetrical faults, however, the rotor voltage
also has a large and permanent negative-sequence component
[see (46)], and the rotor voltages are higher and more damaging
than those for symmetrical grid dips [191].
Table III shows the per unit (p.u.) values of the positivesequence, negative-sequence, and natural fluxes in (44) as a
function of the grid fault type and the depth of the voltage sag
d in p.u. (i.e., for a three-phase voltage sag where the voltage
falls from 1 to 0.2 p.u., the d value is 0.8). The natural flux
value depends on the time instant within the voltage period
where the fault occurs. The phase-to-phase fault presents the
highest natural and negative-sequence flux, and the highest
overvoltages in the rotor windings [191].
The maximum amplitude of the transient rotor voltage is
given in (46) for a symmetrical fault [182]. A discussion of the
maximum rotor voltage amplitude produced by an asymmetrical fault [see (47)] is reported in [99]. As shown in Table III and
(44), deeper voltage sags lead to higher transient voltages,
and larger dip asymmetry increases negative-sequence voltages
and the maximum rotor voltages, as shown in the following,
respectively:
(vro_transient )max

L0
(|s|(1 d)v1s + (1 s)dv1s ) (46)
Ls

(vro_asym )max

L0 |se | v1s + |(s 2)e | v2s


. (47)
Ls
e

Without specific control action, the rotor overvoltages produce high ac rotor currents with synchronous frequencies superposed upon the low-frequency steady-state rotor currents
injected by the RSC [156], [182], [192]. The rotor overcurrent
may exceed 23 times the nominal rotor current, which is not
acceptable [192]. On the stator side, these currents appear as dc
components [192], [193].

The higher rotor currents lead to rising dc-link voltage [156],


[192], [194]. If the system in Fig. 1 is assumed, the GSC
controller intends to regulate the dc-link voltage to its nominal
value, causing a GSC overcurrent of up to 1.5 times the nominal
value. Even with GSC control action, the dc-link voltage can
reach values of about 23 times higher than the nominal dc-link
voltage, beyond the limit of the dc-link capacitor [192].
The positive-sequence flux produces a similar torque behavior to that of the balanced operation, but the negative-sequence
flux tends to create motoring action that results in an increase in
torque pulsation at twice the synchronous frequency [89], [99],
[191] and a reduction in the average torque [195]. The presence
of the second harmonic in the electromagnetic torque can cause
undesired mechanical oscillations, reducing the turbine life
span and creating higher acoustic noise [89], [195].
During the grid disturbance, there is a mismatch between
the mechanical and electromagnetic torque that leads to rotor
overspeed [196]. However, this is not too significant since the
rotor inertia acts as a storage system for the energy surplus, and
a certain increase in speed (10%15%) is acceptable [156].
An induction machine fed by unbalanced voltages produces
unbalanced flux [191] that can lead to unexpected magnetic
saturation, excessive heating, and reduced generator lifetime
[195]. Moreover, it will draw unbalanced currents that will
increase the grid-voltage unbalance and cause overcurrent problems [89].
Immediately after the voltage sag clearance, the sudden change
in the stator voltages causes the natural flux [182] to appear again.
This causes the electromagnetic torque to oscillate, causing
increased stress on the turbine shaft [192].
C. Systems and Control for LVRT Compliance With DFIG
As stated earlier, the grid disturbances cause rotor overcurrents and overvoltages together with a dc-link overvoltage that
can lead to converter failure if no protection is included [182],
[191], [192], [197]. Different protection devices are depicted in
Fig. 22. Their operation and some control approaches to comply
with the LVRT requirements will be discussed here.
The initial solution implemented by manufacturers to protect
the rotor and the converter was to short-circuit the rotor windings with the so-called crowbar and to disconnect the turbine
from the grid [198], [199]. This solution is not allowed with
the LVRT requirements set at the current GCRs because the
WECSs do not support the utility to resume normal operation.
If the RSC is sized to generate a voltage equal to the rotor
overvoltages of (46) and (47), it will be able to fully control
the rotor currents [182], [186], [191]. This is the best solution
to deal with rotor overvoltages because it allows full control
of the DFIG at all times. To achieve it, overmodulation in the
RSC would be required [200], in spite of the increased rotor
current harmonics. A method to design the RSC size based on
the maximum rotor overvoltage and overcurrent is presented
in [90]. One of the most extensive analysis of the operation
limits for the RSC under grid disturbance is presented in [201],
which considers the impact of limited ratings for the GSC and
RSC during grid disturbances. Oversized converters allow more
controllability, but the DFIG topology loses its advantages of

CRDENAS et al.: CONTROL SYSTEMS FOR OPERATION OF DFIGs IN WIND ENERGY APPLICATIONS

2791

Fig. 22. Rotor and converter protection devices: crowbar, dc-link chopper, ESS, and ac switch [156], [213].

the low-size power converter [182]. It is well known that the


converter is sized to manage 30% of the total DFIG power
[3], [181], [202] and is not normally rated to generate a voltage
equal to the rotor overvoltages [182]. It is also noted that, for
deep grid-voltage sags, the RSC oversizing is far beyond the
converter steady-state ratings. Converter sizing is thus a tradeoff
between the LVRT requirements and the cost, together with
other protection elements such as the crowbar and the dc-link
chopper.
DFIGs are always equipped with a crowbar, as shown in
Fig. 22, which is a device that short-circuits the rotor windings
through resistors, thereby limiting the rotor voltage and providing an additional path for the rotor current [185]. Two crowbar
options are available [156]. The first option is the passive
crowbar implemented with a diode rectifier or two thyristors in
antiparallel. This implementation requires the crowbar current
to be forced to zero to deactivate the device, and full control
over the crowbar deactivation is not possible. The second option
is the active crowbar using IGBT switches; this allows crowbar
deactivation and, consequently, faster recovery of the DFIG
control. The crowbar resistance value affects the rotor and
current behavior [197]. Large crowbars result in better damping
of the rotor and stator overcurrents, and the torque overshoot.
It also reduces the reactive power consumption. However, very
large crowbars can cause current spikes upon deactivation and a
high voltage at the rotor slip rings, resulting in voltage stress on
the rotor windings [193], [194]. In [194], Kasem et al. suggest
a crowbar resistance of 0.3 p.u. if the maximum rotor voltage
is limited to 1.2 p.u. The calculation of the crowbar resistance
using

2(vr )max e Ls


(48)
(Rcrowbar )max = 
3.2Vs2 2(vr )2max
is discussed in [193], where (vr )max is the maximum allowable
rotor voltage, Ls = Ls + Lr L0 /(Lr + L0 ), and Vs is the stator
voltage.
Upon activation of the crowbar, the RSC can be switched
off [192], [194], [203]. However, the rotor currents continue

to circulate to the converter dc-link through the freewheeling


diodes of the RSC, leading to a very fast dc-link voltage
increase and a possible activation of the dc-link chopper to limit
the dc-link voltage value [192], [203].
During crowbar operation, rotor currents are not controlled
by the RSC, and the machine acts as a single-fed induction
generator with rotor resistors. The machine consumes reactive
power that can contribute to deepening the grid-voltage sag
[203]. The GSC must supply the grid with reactive power, as
demanded by the LVRT requirements, and the reactive power
to the machine [204], [205]. In [194], it is proposed to connect
the GSC and the RSC in parallel, using suitable ac switches, to
supply more reactive power to the grid.
If the DFIG is not able to supply the reactive power support required by the GCR, dynamic VAR compensators, static
VAR compensators [206], or static synchronous compensators
[207][210] can be installed at the DFIG terminals to provide
it. Other equipment, such as the dynamic voltage restorer, can
also be used [211].
After the fault clearance, transient rotor overvoltages appear
again, and the system experiences a disturbance similar to that
of the initial fault. This would require a crowbar [or dc-link
chopper activation (see Fig. 22)] for a second time [192], [194].
Unlike asymmetrical disturbances, symmetrical grid disturbances only cause transient rotor overvoltages, and the crowbar
mode is active until the rotor currents die down. After this,
the crowbar is disconnected, and the RSC is started again to
control the rotor currents. Since the fault is still present, the
active power reference is reduced to avoid overload. The DFIG
can contribute to the reactive power support to the grid. Note
however that reactive power support is provided by the GSC
throughout the crowbar mode period [192], [204], [212]. In
[214], the crowbar is disconnected when the rotor currents fall
below a threshold value instead of reaching zero, reducing the
crowbar mode time.
The dc-link chopper [156], as shown in Fig. 22, is another
protective device to keep the dc-link voltage within acceptable
limits. It can concurrently operate with the crowbar [156],
[192], [203]. The dc-link chopper is not essential for fault

2792

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 7, JULY 2013

ride-through operation, but it increases the range of DFIG


operation [192], [203]. The ESS [213], [215] connected to
the dc-link absorbs the extra energy supplied to the dc link
and returns it to the DFIG in normal operation. However, it
significantly increases the complexity and cost of the WECS.
A good performance comparison using a crowbar, a dc-link
chopper, and ESS methods is found in [213].
The stator switch shown in Fig. 22, [156], [185], [216]
is another device to meet the LVRT requirements. The stator is disconnected for a short period using this switch; the
RSC is blocked, and the generator is demagnetized. After the
RSC is restarted, the stator is reconnected, and the operation
is resumed. During stator disconnection, the GSC supplies
reactive power to the grid. This implementation limits the
transient magnitude and duration and keeps full control over
the generator during the largest part of the disturbance interval
[156], [216].
As discussed earlier, there is a mismatch between the electromechanical torque and the mechanical torque in the presence
of the grid disturbance. Pitch control can also be used to reduce
the mechanical torque [156], [192] to avoid rotor overspeed.
However, pitch control can change the blade angle at a relatively slow rate [217], which is too slow to help the system to
respond to a grid fault.

Another control approach introduces a virtual resistance in


the rotor to reduce rotor overcurrents. A combination of demagnetization and virtual resistance control is found in [188]. For
symmetrical dips, reduced rotor currents in comparison with
[186] are reported. Operation under asymmetrical faults has not
been reported.
A PI controller with a resonant compensator is presented
in [110], [224], and [225] for operation under distorted gridvoltage conditions. Although results seem promising, the LVRT
issue is not addressed.
In [226] and [227], the conventional controller used in normal operation is switched to a vector-based hysteresis current controller during grid faults. Good system performance
is achieved; however, the operation limits are not specified,
and there are drawbacks to the hysteresis control: higher harmonic content, higher switching frequency or, if the maximum
switching frequency is limited, large error bands that produce
significant low-order harmonics.
Sliding control has been successfully applied to DFIG in
[228] under unbalanced conditions and a harmonically distorted
grid. Future application of this control method to the LVRT
problem can be expected.

VII. C ONCLUSION
D. Control Methods for LVRT Compliance With DFIG
This subsection summarizes the control methods for LVRT
compliance. The goal is to control rotor voltages and currents,
to reduce the rotor overvoltages and/or overcurrents, and to
avoid the crowbar activation in order to keep full DFIG control
at all times to meet the LVRT requirements. However, in
many cases, the crowbar activation cannot be avoided, and the
crowbar mode concurrently works with the control method.
Some control approaches regulate rotor and GSC currents
in the positive and negative dq reference frames [90], [98],
[219][221] based on a positive- and negative-sequence models
of the DFIG [99]. The main control goals cover the DFIG
active and reactive power to meet the LVRT requirements.
As discussed in Section III-B, each power converter has four
degrees of freedom, allowing to include additional control goals
as, for instance, the regulation of the dc-link voltage, stator
current balancing, and cancelation of the oscillations in the
active power, rotor current, and torque.
Although crowbar activation cannot be avoided in case of
severe asymmetrical faults [90], a noncrowbar method to reduce
the rotor overvoltages based on injecting demagnetizing flux
currents from the RSC is proposed in [33], [186], [221], and
[222]. Full DFIG control is retained, but a large rotor current
capacity is needed, and there is limited capability in the case of
asymmetrical faults. If the crowbar is activated, the use of the
demagnetizing current reduces the crowbar mode time [223].
A robust controller in the stationary frame is presented
in [106], claiming full control in all LVRT cases. However, the
results have been obtained with an oversized converter that can
accommodate rotor overvoltages and full rotor current control.
With a suitable-sized converter, this control method may have
some limitations.

This paper has summarized the most recent research in


the field of control systems for DFIGs in wind energy applications. After reviewing the papers related to conventional
control methods for DFIGs connected to balanced systems,
it is concluded that vector control, typically orientated along
the stator flux, is still the most adopted method for regulating
the rotor currents of DFIGs. With this control methodology,
decoupling of the reactive power and electrical torque is simple
to achieve. However, as discussed in Section II, most of the
control schemes presented in Section II-DF can provide good
overall performance.
Regarding sensorless control of variable-speed DFIGs, the
most popular methods are based in MRAS schemes, with the
RCMO providing good performance in both stand-alone and
grid-connected operation of DFIGs. The TBMO is also an
interesting method for sensorless vector control, particularly
because the direct and quadrature rotor currents can be directly
obtained from the components of the signals without
resorting to transformations to a synchronous rotating axis.
Concerning sensorless methods, more research can be required
in some areas, particularly because the performance of the rotor
position observers proposed in the literature have not been
evaluated for LVRT operation.
In this paper, the control systems for the operation of DFIGs
connected to unbalanced grid or loads, have also been assessed.
Several control targets for unbalanced operation have been
proposed in the literature, e.g., to eliminate the oscillations
in the total active power output from the DFIG, to reduce
the oscillations in the total reactive power supplied to the
network, or to supply a grid current with no negative-sequence
components. To fulfill these control targets, the RSC and/or
the GSC can be used. The current trend is to use both power

CRDENAS et al.: CONTROL SYSTEMS FOR OPERATION OF DFIGs IN WIND ENERGY APPLICATIONS

converters simultaneously because more degrees of freedoms


are available in this case.
Control systems for ancillary services and grid-frequency
support have also been discussed in this paper. In the past,
DFIGs where mostly controlled for MPPT operation. Nowadays, it is expected that WECSs based on DFIGs can provide
droop control and inertia emulation. This has been reviewed in
this paper.
Finally, in this paper, LVRT control systems for DFIGs have
been discussed. The operation of the elements typically used for
LVRT compliance, such as crowbars, choppers, static switches,
and other elements, has been analyzed and extensively discussed in this paper.

R EFERENCES
[1] W. Leonhard, Control of Electrical Drives. New York, NY, USA:
Springer-Verlag, 2001.
[2] B. Bose, Modern Power Electronics and AC Drives. Englewood Cliffs,
NJ, USA: Prentice-Hall, 2002.
[3] M. Liserre, R. Cardenas, M. Molinas, and J. Rodriguez, Overview of
multi-MW wind turbines and wind parks, IEEE Trans. Ind. Electron.,
vol. 58, no. 4, pp. 10811095, Apr. 2011.
[4] Web site The wind Power Net, Accessed on May 2012. [Online].
Available: http://www.thewindpower.net
[5] L. H. Hansen, L. Helle, F. Blaabjerg, E. Ritchie, and S. Munk-Nielsen,
Conceptual survey of Generators and Power Electronics for Wind Turbines, Ris Nat. Lab., Roskilde, Denmark, Tech. Rep., 2001.
[6] H. Li, Z. Chen, and H. Polinder, Optimization of multibrid permanentmagnet wind generator systems, IEEE Trans. Energy Convers., vol. 24,
no. 1, pp. 8292, Mar. 2009.
[7] H. Polinder, F. F. A. van der Pijl, G.-J. de Vilder, and P. J. Tavner,
Comparison of direct-drive and geared generator concepts for wind
turbines, IEEE Trans. Energy Convers., vol. 21, no. 3, pp. 725733,
Sep. 2006.
[8] H. Polinder, F. F. A. van der Pijl, G.-J. de Vilder, and P. Tavner, Comparison of direct-drive and geared generator concepts for wind turbines,
in Proc. IEEE Int. Conf. Elect. Mach. Drives, 2005, pp. 543550.
[9] Enercon. (2010). ENERCON wind energy converters Technology &
Service, Aurich, Germany. [Online]. Available: http://www.enercon.de/
p/downloads/EN_Eng_TandS_0710.pdf
[10] V. Delli Colli, F. Marignetti, and C. Attaianese, Analytical and multiphysics approach to the optimal design of a 10-MW DFIG for directdrive wind turbines, IEEE Trans. Ind. Electron., vol. 59, no. 7,
pp. 27912799, Jul. 2012.
[11] C. Xia, Q. Geng, X. Gu, T. Shi, and Z. Song, Inputoutput feedback linearization and speed control of a surface permanent-magnet synchronous
wind generator with the boost-chopper converter, IEEE Trans. Ind.
Electron., vol. 59, no. 9, pp. 34893500, Sep. 2012.
[12] J. M. Espi and J. Castello, Wind turbine generation system with optimized DC-link design and control, IEEE Trans. Ind. Electron., vol. 60,
no. 3, pp. 919929, Mar. 2013.
[13] Vestas, Last Accesed on Sep. 2012. [Online]. Available: http://www.
vestas.com/en/wind-power-plants/procurement/turbine-overview.
aspx#/vestas-univers
[14] Clipper, Last Accessed on September 2012. [Online]. Available: http://
www.clipperwind.com/productline.html
[15] R. Pena, J. C. Clare, and G. M. Asher, Doubly fed induction generator
using back-to-back PWM converters and its application to variable-speed
wind-energy generation, Proc. Inst. Elect. Eng.Elect. Power Appl.,
vol. 143, no. 3, pp. 231241, May 1996.
[16] R. Pena, J. C. Clare, and G. M. Asher, A doubly fed induction generator
using back-to-back PWM converters supplying an isolated load from
a variable speed wind turbine, Proc. Inst. Elect. Eng.Elect. Power
Appl., vol. 143, no. 5, pp. 380387, Sep. 1996.
[17] S. Muller, M. Deicke, and R. W. De Doncker, Doubly fed induction
generator systems for wind turbines, IEEE Ind. Appl. Mag., vol. 8, no. 3,
pp. 2633, May/Jun. 2002.
[18] F. Barati, H. Oraee, E. Abdi, S. Shao, and R. McMahon, The brushless
doubly-fed machine vector model in the rotor flux oriented reference
frame, in Proc. 34th Annu. IEEE IECON, 2008, pp. 14151420.

2793

[19] M. Boger and A. Wallace, Performance capability analysis of the brushless doubly-fed machine as a wind generator, in Proc. 7th Int. Conf.
Elect. Mach. Drives (Conf. Publ. No. 412), 1995, pp. 458461.
[20] L. Xu, B. Guan, H. Liu, L. Gao, and K. Tsai, Design and control of a
high-efficiency Doubly-Fed Brushless machine for wind power generator application, in Proc. IEEE ECCE, 2010, pp. 24092416.
[21] S. Tohidi, H. Oraee, M. R. Zolghadri, S. Shao, and P. Tavner, Analysis
and enhancement of low-voltage ride-through capability of brushless
doubly fed induction generator, IEEE Trans. Ind. Electron., vol. 60,
no. 3, pp. 11461155, Mar. 2013.
[22] M. Ruviaro, F. Runcos, N. Sadowski, and I. M. Borges, Analysis and
test results of a brushless doubly fed induction machine with rotary
transformer, IEEE Trans. Ind. Electron., vol. 59, no. 6, pp. 26702677,
Jun. 2012.
[23] H. T. Jadhav and R. Roy, A critical review on the grid integration issues
of DFIG based wind farms, in Proc. 10th Int. Conf. Environ. Elect. Eng.,
2011, pp. 14.
[24] M. Tazil, V. Kumar, R. C. Bansal, S. Kong, Z. Y. Dong, and W. Freitas,
Three-phase doubly fed induction generators: An overview, IET Elect.
Power Appl., vol. 4, no. 2, pp. 7589, Feb. 2010.
[25] M. S. Marhaba, S. Farhangi, and M. A. Paymani, Comparison of power
fluctuation between OptiSlip and DFIG controlled wind turbines, in
Proc. 3rd PEDSTC, 2012, pp. 290295.
[26] L. Lin, F. Sun, Y. Yang, and Q. Li, Comparison of reactive power
compensation strategy of wind farm based on OptiSlip wind turbines,
in Proc. Int. Conf. SUPERGEN Supply, 2009, pp. 16.
[27] M. R. Khadraoui and M. Elleuch, Comparison between OptiSlip and
fixed speed wind energy conversion systems, in Proc. 5th Int. MultiConf. SSD, 2008, pp. 16.
[28] V. Inc., Vestas V90-1.8MW and V90-2.0MW, product brochure.
[Online]. Available: http://www.vestas.com/en/media/brochures.aspx
[29] G. Abad, M. A. Rodriguez, G. Iwanski, and J. Poza, Direct power
control of doubly-fed-induction-generator-based wind turbines under
unbalanced grid voltage, IEEE Trans. Power Electron., vol. 25, no. 2,
pp. 442452, Feb. 2010.
[30] H. Akagi and H. Sato, Control and performance of a doubly-fed induction machine intended for a flywheel energy storage system, IEEE
Trans. Power Electron., vol. 17, no. 1, pp. 109116, Jan. 2002.
[31] J. Arbi, M. J.-B. Ghorbal, I. Slama-Belkhodja, and L. Charaabi, Direct virtual torque control for doubly fed induction generator grid connection, IEEE Trans. Ind. Electron., vol. 56, no. 10, pp. 41634173,
Oct. 2009.
[32] I. Cadirci and M. Ermis, Double-output induction generator operating at subsynchronous and supersynchronous speeds: Steady-state performance optimisation and wind-energy recovery, Proc. Inst. Elect.
Eng.Elect. Power Appl., vol. 139, no. 5, pp. 429442, Sep. 1992.
[33] V. Flores Mendes, C. V. de Sousa, S. R. Silva, J. Rabelo, and
W. Hofmann, Modeling and ride-through control of doubly fed induction generators during symmetrical voltage sags, IEEE Trans. Energy
Convers., vol. 26, no. 4, pp. 11611171, Dec. 2011.
[34] P. G. Holmes and N. A. Elsonbaty, Cycloconvertor-excited dividedwinding doubly-fed machine as a wind-power convertor, Proc.
Inst. Elect. Eng.Elect. Power Appl., vol. 131, no. 2, pp. 6169,
Mar. 1984.
[35] J. Hu, H. Nian, B. Hu, Y. He, and Z. Q. Zhu, Direct active and
reactive power regulation of DFIG using sliding-mode control approach, IEEE Trans. Energy Convers., vol. 25, no. 4, pp. 10281039,
Dec. 2010.
[36] J. Hu, J. Zhu, Q. Ma, and Y. Zhang, Predictive direct virtual torque
control of doubly fed induction generator for grid synchronization, in
Proc. ICEMS, 2011, pp. 16.
[37] S. Li, T. A. Haskew, K. A. Williams, and R. P. Swatloski, Control
of DFIG wind turbine with direct-current vector control configuration,
IEEE Trans. Sustain. Energy, vol. 3, no. 1, pp. 111, Jan. 2012.
[38] Z. Liu, O. A. Mohammed, and S. Liu, A novel direct torque control
of doubly-fed induction generator used for variable speed wind power
generation, in Proc. IEEE Power Eng. Soc. Gen. Meeting, 2007, pp. 16.
[39] B. B. Pimple, V. Y. Vekhande, and B. G. Fernandes, A new direct
torque control of doubly-fed induction generator under unbalanced grid
voltage, in Proc. 6th Annu. IEEE APEC Expo., 2011, pp. 15761581.
[40] L. Qizhong, Y. Lan, and W. Guoxiang, Comparison of control strategy
for Double-Fed Induction Generator (DFIG), in Proc. 3rd ICMTMA,
2011, vol. 1, pp. 741744.
[41] B. C. Rabelo, W. Hofmann, J. L. da Silva, R. G. de Oliveira, and
S. R. Silva, Reactive power control design in doubly fed induction
generators for wind turbines, IEEE Trans. Ind. Electron., vol. 56, no. 10,
pp. 41544162, Oct. 2009.

2794

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 7, JULY 2013

[42] D. Santos-Martin, J. L. Rodriguez-Amenedo, and S. Arnalte, Direct


power control applied to doubly fed induction generator under unbalanced grid voltage conditions, IEEE Trans. Power Electron., vol. 23,
no. 5, pp. 23282336, Sep. 2008.
[43] G. A. Smith, Static Scherbius system of induction-motor speed control, Proc. Inst. Elect. Eng., vol. 124, no. 6, pp. 557560, Jun. 1977.
[44] G. A. Smith and K. A. Nigim, Wind-energy recovery by a static
Scherbius induction generator, Proc. Inst. Elect. Eng.Gen., Transmiss. Distrib., vol. 128, no. 6, pp. 317324, Nov. 1981.
[45] E. Tremblay, S. Atayde, and A. Chandra, Comparative study
of control strategies for the doubly fed induction generator in
wind energy conversion systems: A DSP-based implementation approach, IEEE Trans. Sustain. Energy, vol. 2, no. 3, pp. 288299,
Jul. 2011.
[46] M. Y. Uctug, I. Eskandarzadeh, and H. Ince, Modelling and output
power optimisation of a wind turbine driven double output induction
generator, Proc. Inst. Elect. Eng.Elect. Power Appl., vol. 141, no. 2,
pp. 3338, Mar. 1994.
[47] K. C. Wong, S. L. Ho, and K. W. E. Cheng, Direct control algorithm for
doubly fed induction generators in weak grids, IET Elect. Power Appl.,
vol. 3, no. 4, pp. 371380, Jul. 2009.
[48] G. Xiao-Ming, S. Dan, H. Ben-Teng, and H. Ling-Ling, Direct power
control for wind-turbine driven doubly-fed induction generator with constant switch frequency, in Proc. ICEMS, 2007, pp. 253258.
[49] L. Xu and P. Cartwright, Direct active and reactive power control
of DFIG for wind energy generation, IEEE Trans. Energy Convers.,
vol. 21, no. 3, pp. 750758, Sep. 2006.
[50] M. Yamamoto and O. Motoyoshi, Active and reactive power control
for doubly-fed wound rotor induction generator, IEEE Trans. Power
Electron., vol. 6, no. 4, pp. 624629, Oct. 1991.
[51] D. Zhi and L. Xu, Direct power control of DFIG with constant switching
frequency and improved transient performance, IEEE Trans. Energy
Convers., vol. 22, no. 1, pp. 110118, Mar. 2007.
[52] C. Kramer, Neue methode zur regelung von asynchronen motoren,
Elektrotech.Z., vol. 29, p. 734, 1908.
[53] A. Lavi and R. J. Polge, Induction motor speed control with static
inverter in the rotor, IEEE Trans. Power App. Syst., vol. PAS-85, no. 1,
pp. 7684, Jan. 1966.
[54] M. Machmoum, M. Cherkaoui, R. le Doeuff, and F. M. Sargos, Steady
state analysis and experimental investigation of a doubly-fed induction
machine supplied by a current-source cycloconverter in the rotor,
in Proc. 4th Int. Conf. Power Electron. Variable-Speed Drives, 1990,
pp. 8186.
[55] A. Tapia, G. Tapia, J. X. Ostolaza, and J. R. Saenz, Modeling and
control of a wind turbine driven doubly fed induction generator, IEEE
Trans. Energy Convers., vol. 18, no. 2, pp. 194204, Jun. 2003.
[56] R. Cardenas, R. Pena, G. Tobar, J. Clare, P. Wheeler, and G. Asher,
Stability analysis of a wind energy conversion system based on a doubly
fed induction generator fed by a matrix converter, IEEE Trans. Ind.
Electron., vol. 56, no. 10, pp. 41944206, Oct. 2009.
[57] R. Pena, R. Cardenas, E. Reyes, J. Clare, and P. Wheeler, A topology
for multiple generation system with doubly fed induction machines and
indirect matrix converter, IEEE Trans. Ind. Electron., vol. 56, no. 10,
pp. 41814193, Oct. 2009.
[58] J. A. Santisteban and R. M. Stephan, Vector control methods for induction machines: An overview, IEEE Trans. Educ., vol. 44, no. 2, pp. 170
175, May 2001.
[59] G. Escobar, M. F. Martinez-Montejano, A. A. Valdez, P. R. Martinez,
and M. Hernandez-Gomez, Fixed-reference-frame phase-locked loop
for grid synchronization under unbalanced operation, IEEE Trans. Ind.
Electron., vol. 58, no. 5, pp. 19431951, May 2011.
[60] R. Pena, R. Cardenas, J. Proboste, J. Clare, and G. Asher, Wind-diesel
generation using doubly fed induction machines, IEEE Trans. Energy
Convers., vol. 23, no. 1, pp. 202214, Mar. 2008.
[61] R. Pea, R. Cardenas, J. Proboste, G. Asher, and J. Clare, Sensorless
control of doubly-fed induction generators using a rotor-current-based
MRAS observer, IEEE Trans. Ind. Electron., vol. 55, no. 1, pp. 330
339, Jan. 2008.
[62] B. Rabelo and W. Hofmann, Power flow optimisation and grid integration of wind turbines with the doubly-fed induction generator, in Proc.
IEEE Power Electron. Spec. Conf., 2005, pp. 29302936.
[63] B. Andresen and J. Birk, A high power density converter system for
the Gamesa G10x 4, 5 MW wind turbine keywords, in Proc. Eur. Conf.
Power Electron. Appl., 2007, pp. 18.
[64] T. Zhang and A. Zain, Modular converter system reliability & performance analysis in design, in Proc. 2nd IEEE Int. Symp. Power Electron.
Distrib. Gen. Syst., Jun. 2010, pp. 252258.

[65] J. Llorente, A. Bjorn, and J. Birk, Method for operation of a converter


system, Gamesa Patent 754 505 2B2, May 14, 2008.
[66] N. R. N. Idris, C. L. Toh, and M. E. Elbuluk, A new torque and flux
controller for direct torque control of induction machines, IEEE Trans.
Ind. Appl., vol. 42, no. 6, pp. 13581366, Nov./Dec. 2006.
[67] F. Bonnet, P.-E. Vidal, and M. Pietrzak-David, Dual direct torque
control of doubly fed induction machine, IEEE Trans. Ind. Electron.,
vol. 54, no. 5, pp. 24822490, Oct. 2007.
[68] G. Abad, M. A. Rodriguez, J. Poza, and J. M. Canales, Direct torque
control for doubly fed induction machine-based wind turbines under
voltage dips and without crowbar protection, IEEE Trans. Energy Convers., vol. 25, no. 2, pp. 586588, Jun. 2010.
[69] G. Abad, M. A. Rodriguez, and J. Poza, Two-level VSC based predictive direct torque control of the doubly fed induction machine with
reduced torque and flux ripples at low constant switching frequency,
IEEE Trans. Power Electron., vol. 23, no. 3, pp. 10501061, May 2008.
[70] ABB, ABB low voltage wind turbine converters ACS800-67,
0.6 to 2.2 MW, Zurich, Switzerland, 2012. [Online]. Available:
http://www05.abb.com/global/scot/scot201.nsf/veritydisplay/
d971c93658b48422c1257799002541f5/$file/acs800-67_wind%
20turbine%20converter_flyer_3aua0000086088_en_rev%20a.pdf
[71] T. Noguchi, H. Tomiki, S. Kondo, and I. Takahashi, Direct power
control of PWM converter without power-source voltage sensors, IEEE
Trans. Ind. Appl., vol. 34, no. 3, pp. 473479, May/Jun. 1998.
[72] G. Escobar, A. M. Stankovic, J. M. Carrasco, E. Galvan, and R. Ortega,
Analysis and design of Direct Power Control (DPC) for a three phase
synchronous rectifier via output regulation subspaces, IEEE Trans.
Power Electron., vol. 18, no. 3, pp. 823830, May 2003.
[73] M. Malinowski, M. P. Kazmierkowski, S. Hansen, F. Blaabjerg, and
G. D. Marques, Virtual-flux-based direct power control of three-phase
PWM rectifiers, IEEE Trans. Ind. Appl., vol. 37, no. 4, pp. 10191027,
Jul./Aug. 2001.
[74] R. Datta and V. T. Ranganathan, Direct power control of grid-connected
wound rotor induction machine without rotor position sensors, IEEE
Trans. Power Electron., vol. 16, no. 3, pp. 390399, May 2001.
[75] R. Cardenas, R. Pena, M. Perez, J. Clare, G. Asher, and P. Wheeler,
Control of a switched reluctance generator for variable-speed wind energy applications, IEEE Trans. Energy Convers., vol. 20, no. 4, pp. 781
791, Dec. 2005.
[76] E. Echenique, J. Dixon, R. Cardenas, and R. Pena, Sensorless control
for a switched reluctance wind generator, based on current slopes and
neural networks, IEEE Trans. Ind. Electron., vol. 56, no. 3, pp. 817
825, Mar. 2009.
[77] M. A. Abdullah, A. H. M. Yatim, C. W. Tan, and R. Saidur, A review
of maximum power point tracking algorithms for wind energy systems,
Renew. Sustain. Energy Rev., vol. 16, no. 5, pp. 32203227, Jun. 2012.
[78] R. Cardenas and R. Pena, Sensorless vector control of induction machines for variable-speed wind energy applications, IEEE Trans. Energy
Convers., vol. 19, no. 1, pp. 196205, Mar. 2004.
[79] C. Evangelista, P. Puleston, F. Valenciaga, and L. M. Fridman,
Lyapunov-designed super-twisting sliding mode control for wind energy conversion optimization, IEEE Trans. Ind. Electron., vol. 60, no. 2,
pp. 538545, Feb. 2013.
[80] Y. Xia, K. H. Ahmed, and B. W. Williams, Wind turbine power coefficient analysis of a new maximum power point tracking technique, IEEE
Trans. Ind. Electron., vol. 60, no. 3, pp. 11221132, Mar. 2013.
[81] F. Yao, Z. Y. Dong, K. Meng, Z. Xu, H. H. Iu, and K. P. Wong,
Quantum-inspired particle swarm optimization for power system operations considering wind power uncertainty and carbon tax in Australia,
IEEE Trans. Ind. Informat., vol. 8, no. 4, pp. 880888, Nov. 2012.
[82] A. Mirecki, X. Roboam, and F. Richardeau, Comparative study of
maximum power strategy in wind turbines, in Proc. IEEE Int. Symp.
Ind. Electron., 2004, pp. 993998.
[83] L. G. Gonzlez, E. Figueres, G. Garcer, and O. Carranza, Dynamic
response analysis of small wind energy conversion systems (WECS)
operating with torque control versus speed control, in Proc. ICREPQ,
2009, pp. 15.
[84] G. Tapia and A. Tapia, Wind generation optimisation algorithm for a
doubly fed induction generator, Proc. Inst. Elect. Eng.Gen. Transmiss. Distrib., vol. 152, no. 2, pp. 253263, Mar. 2005.
[85] H. Camblong, I. M. de Alegria, M. Rodriguez, and G. Abad, Experimental evaluation of wind turbines maximum power point tracking
controllers, Energy Convers. Manage., vol. 47, no. 18/19, pp. 2846
2858, Nov. 2006.
[86] M. A. Abdullah, A. H. M. Yatim, and C. Wei, A study of maximum
power point tracking algorithms for wind energy system, in Proc. IEEE
Conf. CET, 2011, pp. 321326.

CRDENAS et al.: CONTROL SYSTEMS FOR OPERATION OF DFIGs IN WIND ENERGY APPLICATIONS

[87] C. Patsios, A. Chaniotis, M. Rotas, and A. G. Kladas, A comparison


of maximum-power-point tracking control techniques for low-power
variable-speed wind generators, in Proc. ELECTROMOTION, Jul. 13,
2009, pp. 16.
[88] T. Brekken, N. Mohan, and T. Undeland, Control of a doubly-fed induction wind generator under unbalanced grid voltage conditions, in Proc.
Eur. Conf. Power Electron. Appl., 2005, p. 10.
[89] T. K. A. Brekken and N. Mohan, Control of a doubly fed induction
wind generator under unbalanced grid voltage conditions, IEEE Trans.
Energy Convers., vol. 22, no. 1, pp. 129135, Mar. 2007.
[90] M. I. Martinez, G. Tapia, A. Susperregui, and H. Camblong, DFIG
power generation capability and feasibility regions under unbalanced
grid voltage conditions, IEEE Trans. Energy Convers., vol. 26, no. 4,
pp. 10511062, Dec. 2011.
[91] R. Pea, R. Crdenas, E. Escobar, J. Clare, and P. Wheeler, Control
system for unbalanced operation of stand-alone doubly fed induction
generators, IEEE Trans. Energy Convers., vol. 22, no. 2, pp. 544545,
Jun. 2007.
[92] R. Pena, R. Cardenas, and E. Escobar, Control strategy for a doublyfed induction generator feeding an unbalanced grid or stand-alone load,
Elect. Power Syst. Res., vol. 79, no. 2, pp. 355364, Feb. 2009.
[93] V.-T. Phan and H.-H. Lee, Improved predictive current control for unbalanced stand-alone doubly-fed induction generator-based wind power
systems, IET Elect. Power Appl., vol. 5, no. 3, pp. 275287, Mar. 2011.
[94] G. Iwanski and W. Koczara, Sensorless direct voltage control of the
stand-alone slip-ring induction generator, IEEE Trans. Ind. Electron.,
vol. 54, no. 2, pp. 12371239, Apr. 2007.
[95] V.-T. Phan and H.-H. Lee, Performance enhancement of stand-alone
DFIG systems with control of rotor and load side converters using resonant controllers, IEEE Trans. Ind. Appl., vol. 48, no. 1, pp. 199210,
Jan./Feb. 2012.
[96] G. H. Ghorashi, S. Murthy, B. P. Singh, and B. Singh, Analysis of
wind driven grid connected induction generators under unbalanced grid
conditions, IEEE Trans. Energy Convers., vol. 9, no. 2, pp. 217223,
Jun. 1994.
[97] E. Muljadi, T. Batan, D. Yildirim, and C. P. Butterfield, Understanding
the unbalancedvoltage problem in wind turbine generation, in Conf.
Rec. IEEE IAS Annu. Meeting, Phoenix, AZ, USA, 1999, pp. 13591365.
[98] Y. Zhou, P. Bauer, J. A. Ferreira, and J. Pierik, Operation of gridconnected DFIG under unbalanced grid voltage condition, IEEE Trans.
Energy Convers., vol. 24, no. 1, pp. 240246, Mar. 2009.
[99] L. Xu and Y. Wang, Dynamic modeling and control of DFIG-based
wind turbines under unbalanced network conditions, IEEE Trans.
Power Syst., vol. 22, no. 1, pp. 314323, Feb. 2007.
[100] T. Brekken and N. Mohan, A novel doubly-fed induction wind generator
control scheme for reactive power control and torque pulsation compensation under unbalanced grid voltage conditions, in Proc. 34th IEEE
Annu. Power Electron. Spec. Conf., Acapulco, Mexico, 2003, pp. 1519.
[101] Y. Wang, L. Xu, and B. W. Williams, Compensation of network voltage
unbalance using doubly fed induction generator-based wind farms, IET
Renew. Power Gen., vol. 3, no. 1, pp. 1222, Mar. 2009.
[102] J. Hu, Y. He, L. Xu, and B. W. Williams, Improved control of DFIG
systems during network unbalance using PI-R current regulators, IEEE
Trans. Ind. Electron., vol. 56, no. 2, pp. 439451, Feb. 2009.
[103] J. Hu and Y. He, Reinforced control and operation of DFIG-based windpower-generation system under unbalanced grid voltage conditions,
IEEE Trans. Energy Convers., vol. 24, no. 4, pp. 905915, Dec. 2009.
[104] V.-T. Phan, S.-H. Kwak, and H.-H. Lee, An improved control method
for DFIG-based wind system supplying unbalanced stand-alone loads,
in Proc. IEEE ISIE, 2009, pp. 10811086.
[105] V.-T. Phan and H.-H. Lee, Control of an unbalanced stand-alone DFIGbased wind system using predictive current control method, in Proc.
IEEE ECCE, 2010, pp. 25212528.
[106] J. P. da Costa, H. Pinheiro, T. Degner, and G. Arnold, Robust controller
for DFIGs of grid-connected wind turbines, IEEE Trans. Ind. Electron.,
vol. 58, no. 9, pp. 40234038, Sep. 2011.
[107] B. Singh and S. Sharma, Design and implementation of four-leg
voltage-source-converter-based VFC for autonomous wind energy conversion system, IEEE Trans. Ind. Electron., vol. 59, no. 12, pp. 4694
4703, Dec. 2012.
[108] R. Cardenas, C. Juri, R. Pena, J. Clare, and P. Wheeler, Analysis and
experimental validation of control systems for four-leg matrix converter
applications, IEEE Trans. Ind. Electron., vol. 59, no. 1, pp. 141153,
Jan. 2012.
[109] R. Cardenas, R. Pena, P. Wheeler, and J. Clare, Experimental validation
of a space vector modulation algorithm for four-leg matrix converters,
IEEE Trans. Ind. Electron., vol. 58, no. 4, pp. 12821293, Apr. 2011.

2795

[110] H. Xu, J. Hu, and Y. He, Operation of wind-turbine-driven DFIG systems under distorted grid voltage conditions: Analysis and experimental
validations, IEEE Trans. Power Electron., vol. 27, no. 5, pp. 23542366,
May 2012.
[111] Y. Wang and L. Xu, Coordinated control of DFIG and FSIG-based
wind farms under unbalanced grid conditions, IEEE Trans. Power Del.,
vol. 25, no. 1, pp. 367377, Jan. 2010.
[112] P. Zhou, Y. He, and D. Sun, Improved direct power control of a DFIGbased wind turbine during network unbalance, IEEE Trans. Power
Electron., vol. 24, no. 11, pp. 24652474, Nov. 2009.
[113] R. Pena, R. Cardenas, G. M. Asher, J. C. Clare, J. Rodriguez, and
P. Cortes, Vector control of a diesel-driven doubly fed induction machine for a stand-alone variable speed energy system, in Proc. 28th
Annu. IEEE IECON, 2002, vol. 2, pp. 985990.
[114] R. Pena, R. Cardenas, J. Clare, and G. Asher, Control strategy of doubly
fed induction generators for a wind diesel energy system, in Proc. 28th
Annu. IEEE IECON, 2002, vol. 4, pp. 32973302.
[115] R. Cardenas, R. Pena, J. Proboste, G. Asher, and J. Clare, MRAS
observer for sensorless control of standalone doubly fed induction generators, IEEE Trans. Energy Convers., vol. 20, no. 4, pp. 710718,
Dec. 2005.
[116] A. Griffo, D. Drury, T. Sawata, and P. H. Mellor, Sensorless starting of
a wound-field synchronous starter/generator for aerospace applications,
IEEE Trans. Ind. Electron., vol. 59, no. 9, pp. 35793587, Sep. 2012.
[117] M. Abolhassani, P. Niazi, H. Tolivat, and P. Enjeti, A sensorless Integrated Doubly-Fed Electric Alternator/Active filter (IDEA) for variable
speed wind energy system, in Conf. Rec. 38th IEEE IAS Annu. Meeting,
2003, pp. 507514.
[118] R. Datta and V. T. Ranganathan, A simple position-sensorless algorithm
for rotor-side field-oriented control of wound-rotor induction machine,
IEEE Trans. Ind. Electron., vol. 48, no. 4, pp. 786793, Aug. 2001.
[119] L. Morel, H. Godfroid, A. Mirzaian, and J. M. Kauffmann, Doublefed induction machine: Converter optimisation and field oriented control
without position sensor, Proc. Inst. Elect. Eng.Elect. Power Appl.,
vol. 145, no. 4, pp. 360368, Jul. 1998.
[120] E. Bogalecka and Z. Krzeminski, Sensorless control of a double-fed
machine for wind power generators, in Proc. EPE-PMC, Dubrovnik
and Cavtat, Croatia, 2002, [CD-ROM].
[121] B. Hopfensperger, D. J. Atkinson, and R. A. Lakin, Stator-flux-oriented
control of a doubly-fed induction machine with and without position
encoder, Proc. Inst. Elect. Eng.Elect. Power Appl., vol. 147, no. 4,
pp. 241250, Jul. 2000.
[122] O. A. Mohammed, Z. Liu, and S. Liu, A novel sensorless control strategy of doubly fed induction motor and its examination with the physical
modeling of machines, IEEE Trans. Magn., vol. 41, no. 5, pp. 1852
1855, May 2005.
[123] L. Xu and W. Cheng, Torque and reactive power control of a doubly
fed induction machine by position sensorless scheme, IEEE Trans. Ind.
Appl., vol. 31, no. 3, pp. 636642, May/Jun. 1995.
[124] C. Schauder, Adaptive speed identification for vector control of induction motors without rotational transducers, IEEE Trans. Ind. Appl.,
vol. 28, no. 5, pp. 10541061, Sep./Oct. 1992.
[125] R. Ghosn, C. Asmar, M. Pietrzak-David, and B. De Fornel, A MRASsensorless speed control of doubly fed induction machine, in Proc. Int.
Conf. Elect. Mach., 2002, pp. 2628.
[126] R. Ghosn, C. Asmar, M. Pietrzak-David, and B. De Fornel, A MRAS
Luenberger sensorless speed control of doubly fed induction machine,
in Proc. Eur. Power Electron. Conf., 2003, [CD-ROM].
[127] R. Cardenas, R. Pena, G. Asher, J. Clare, and J. Cartes, MRAS observer
for doubly fed induction machines, IEEE Trans. Energy Convers.,
vol. 19, no. 2, pp. 467468, Jun. 2004.
[128] R. Cardenas, R. Pena, J. Proboste, G. Asher, and J. Clare, Rotor current
based MRAS observer for doubly-fed induction machines, Electron.
Lett., vol. 40, no. 12, pp. 769770, Jun. 2004.
[129] M. Pattnaik and D. Kastha, Comparison of MRAS based speed estimation methods for a stand-alone doubly fed induction generator, in Proc.
ICEAS, 2011, pp. 16.
[130] Y. Guofeng, L. Yongdong, C. Jianyun, and J. Xinjian, A novel position
sensor-less control scheme of Doubly Fed Induction Wind Generator
based on MRAS method, in Proc. IEEE PESC, 2008, pp. 27232727.
[131] M. S. Carmeli, M. Iacchetti, and R. Perini, A MRAS observer applied
to sensorless doubly fed induction machine drives, in Proc. IEEE ISIE,
2010, vol. 1, pp. 30773082.
[132] D. G. Forchetti, G. O. Garca, S. Member, and M. I. Valla, Adaptive
observer for sensorless control of stand-alone doubly fed induction generator, IEEE Trans. Ind. Electron., vol. 56, no. 10, pp. 41744180,
Oct. 2009.

2796

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 7, JULY 2013

[133] R. Cardenas, R. Pena, J. Clare, G. Asher, and J. Proboste, MRAS


observers for sensorless control of doubly-fed induction generators,
IEEE Trans. Power Electron., vol. 23, no. 3, pp. 10751084, May 2008.
[134] R. Cardenas, R. Pena, J. Proboste, G. Asher, J. Clare, and P. Wheeler,
MRAS observers for sensorless control of doubly-fed induction generators, in Proc. 4th IET Conf. PEMD, 2008, pp. 568572.
[135] F. Castelli-Dezza, G. Foglia, M. F. Iacchetti, and R. Perini, An MRAS
observer for sensorless DFIM drives with direct estimation of the torque
and flux rotor current components, IEEE Trans. Power Electron.,
vol. 27, no. 5, pp. 25762584, May 2012.
[136] R. Pena, R. Cardenas, J. Proboste, G. Asher, and J. Clare, Sensorless
control of a slip ring induction generator based on rotor current MRAS
observer, in Proc. 36th IEEE PESC, 2005, pp. 25082513.
[137] D. Forchetti, G. O. Garcia, and M. I. Valla, Sensorless control of
standalone doubly fed induction generator with an adaptive observer,
in Proc. IEEE ISIE, 2008, pp. 24442449.
[138] K. Ohyama, G. Asher, and M. Sumner, Comparative analysis of experimental performance and stability of sensorless induction motor drives,
IEEE Trans. Ind. Electron., vol. 53, no. 1, pp. 178186, Feb. 2005.
[139] M. F. Iacchetti, Adaptive tuning of the stator inductance in a rotorcurrent-based MRAS observer for sensorless doubly fed inductionmachine drives, IEEE Trans. Ind. Electron., vol. 58, no. 10, pp. 4683
4692, Oct. 2011.
[140] M. S. Carmeli, F. Castelli-dezza, M. Iacchetti, and R. Perini, Effects
of mismatched parameters in MRAS sensorless doubly fed induction
machine drives, IEEE Trans. Power Electron., vol. 25, no. 11, pp. 2842
2851, Nov. 2010.
[141] B. Mwinyiwiwa, Y. Zhang, B. Shen, and B.-T. Ooi, Rotor position
phase-locked loop for decoupled P Q control of DFIG for wind power
generation, IEEE Trans. Energy Convers., vol. 24, no. 3, pp. 758765,
Sep. 2009.
[142] B. Shen, B. Mwinyiwiwa, Y. Zhang, and B.-T. Ooi, Sensorless maximum power point tracking of wind by DFIG using rotor Position Phase
Lock Loop (PLL), IEEE Trans. Power Electron., vol. 24, no. 4, pp. 942
951, Apr. 2009.
[143] K. Gogas, G. Joos, B. T. Ooi, Y. Z. Zhang, and B. Mwinyiwiwa, Design
of a robust speed and position sensorless decoupled P-Q controlled
doubly-fed induction generator for variable-speed wind energy applications, in Proc. IEEE Canada EPC, 2007, pp. 6267.
[144] G. D. Marques and D. M. Sousa, Air-gap-power-vector-based sensorless method for DFIG control without flux estimator, IEEE Trans. Ind.
Electron., vol. 58, no. 10, pp. 47174726, Oct. 2011.
[145] G. D. Marques and D. M. Sousa, A DFIG sensorless method for direct
estimation of slip position, in Proc. IEEE Region 8 Int. Conf. Comput.
Technol. Elect. Electron. Eng. (SIBIRCON), 2010, pp. 818823.
[146] G. D. Marques, V. F. Pires, S. Sousa, and D. M. Sousa, A DFIG
sensorless rotor-position detector based on a hysteresis controller, IEEE
Trans. Energy Convers., vol. 26, no. 1, pp. 917, Mar. 2011.
[147] G. D. Marques, V. F. Pires, S. Sousa, and D. M. Sousa, Evaluation of
a DFIG rotor position-sensorless detector based on a hysteresis controller, in Proc. Int. Conf. POWERENG Energy Elect. Drives, 2009,
pp. 113116.
[148] S. Georges, G. Ragi, P.-D. Maria, and D. F. Bernard, A comparison
of sensorless speed estimation for a doubly fed induction machine, in
Proc. Eur. Conf. Power Electron. Appl., 2005, p. 9.
[149] J. Holtz, Sensorless control of induction machines; With or without
signal injection? IEEE Trans. Ind. Electron., vol. 53, no. 1, pp. 730,
Feb. 2005.
[150] C. Luo, H. G. Far, H. Banakar, P. Keung, B. Ooi, and L. Fellow, Estimation of wind penetration as limited by frequency deviation, IEEE Trans.
Energy Convers., vol. 22, no. 3, pp. 783791, Sep. 2007.
[151] H. Banakar, C. Luo, B. T. Ooi, L. Fellow, A. A. G. C. Feedback, and
L. Time, Impacts of wind power minute-to-minute variations on power
system operation, IEEE Trans. Power Syst., vol. 23, no. 1, pp. 150160,
Feb. 2008.
[152] C. Luo, S. Member, B. Ooi, and L. Fellow, Frequency deviation of
thermal power plants due to wind farms, IEEE Trans. Energy Convers.,
vol. 21, no. 3, pp. 708716, Sep. 2006.
[153] Z.-S. Zhang, Y.-Z. Sun, J. Lin, and G.-J. Li, Coordinated frequency
regulation by doubly fed induction generator-based wind power plants,
IET Renew. Power Gen., vol. 6, no. 1, pp. 3847, Jan. 2012.
[154] X. Zhang, H. Li, and Y. Wang, Control of DFIG-based wind farms for
power network frequency support, in Proc. Int. Conf. POWERCON,
2010, pp. 16.
[155] J. Morren, S. W. H. de Haan, W. L. Kling, and J. A. Ferreira, Wind
turbines emulating inertia and supporting primary frequency control,
IEEE Trans. Power Syst., vol. 21, no. 1, pp. 433434, Feb. 2006.

[156] M. Tsili and S. Papathanassiou, A review of grid code technical requirements for wind farms, IET Renew. Power Gen., vol. 3, no. 3, pp. 308
332, Sep. 2009.
[157] Grid Code: High and Extra High Voltage, E. ON Netz GmbH, Bayreuth,
Germany, 2006.
[158] A. Mullane and M. OMalley, The inertial response of inductionmachine-based wind turbines, IEEE Trans. Power Syst., vol. 20, no. 3,
pp. 14961503, Aug. 2005.
[159] M. Fazeli, G. M. Asher, C. Klumpner, and L. Yao, Novel integration
of DFIG-based wind generators within microgrids, IEEE Trans. Energy
Convers., vol. 26, no. 3, pp. 840850, Sep. 2011.
[160] I. A. Gowaid, A. El-Zawawi, and M. El-Gammal, Improved inertia
and frequency support from grid-connected DFIG wind farms, in Proc.
IEEE/PES PSCE, 2011, pp. 19.
[161] M. Kayikci and J. V. Milanovic, Dynamic contribution of DFIG-based
wind plants to system frequency disturbances, IEEE Trans. Power Syst.,
vol. 24, no. 2, pp. 859867, May 2009.
[162] B. Khaki, M. H. Asgari, R. Sirjani, and A. Mozdawar, Contribution of
DFIG wind turbines to system frequency control, in Proc. Int. Conf.
SUPERGEN Supply, 2009, pp. 18.
[163] X. Zhu, Y. Wang, L. Xu, X. Zhang, and H. Li, Virtual inertia control
of DFIG-based wind turbines for dynamic grid frequency support, in
Proc. IET Conf. RPG, 2011, pp. 16.
[164] F. M. Hughes, O. Anaya-lara, N. Jenkins, and G. Strbac, Control of
DFIG-based wind generation for power network support, IEEE Trans.
Power Syst., vol. 20, no. 4, pp. 19581966, Nov. 2005.
[165] D. Gautam, S. Member, L. Goel, R. Ayyanar, and S. Member, Control
strategy to mitigate the impact of reduced inertia due to doubly fed
induction generators on large power systems, IEEE Trans. Power Syst.,
vol. 26, no. 1, pp. 214224, Feb. 2011.
[166] G. Ramtharan, J. B. Ekanayake, and N. Jenkins, Frequency support
from doubly fed induction generator wind turbines, IET Renew. Power
Gen., vol. 1, no. 1, pp. 39, Mar. 2007.
[167] L.-R. Chang-Chien, W.-T. Lin, and Y.-C. Yin, Enhancing frequency
response control by DFIGs in the high wind penetrated power systems,
IEEE Trans. Power Syst., vol. 26, no. 2, pp. 710718, May 2011.
[168] I. D. Margaris, S. A. Papathanassiou, S. Member, N. D. Hatziargyriou,
A. D. Hansen, and P. Srensen, Frequency control in autonomous power
systems with high wind power penetration, IEEE Trans. Sustain. Energy, vol. 3, no. 2, pp. 189199, Apr. 2012.
[169] L.-R. Chang-Chien and Y. Yin, Strategies for operating wind power
in a similar manner of conventional power plant, IEEE Trans. Energy
Convers., vol. 24, no. 4, pp. 926934, Dec. 2009.
[170] S. El Itani and G. Joos, Assessment of inertial potential of variablespeed wind turbines, in Proc. IEEE ECCE, 2011, pp. 851856.
[171] C. Luo, H. Banakar, B. Shen, and B. Ooi, Strategies to smooth wind
power fluctuations of wind turbine generator, IEEE Trans. Energy Convers., vol. 22, no. 2, pp. 341349, Jun. 2007.
[172] M. Fazeli, G. M. Asher, C. Klumpner, L. Yao, and M. Bazargan, Novel
integration of wind generator-energy storage systems within microgrids, IEEE Trans. Smart Grid, vol. 3, no. 2, pp. 728737, Jun. 2012.
[173] M. Fazeli, G. M. Asher, C. Klumpner, S. Bozhko, L. Yao, and
M. Bazargan, Wind turbine-energy storage control system for delivering constant demand power shared by DFIGs through droop characteristics, in Proc. 13th Eur. Conf. EPE Appl., 2009, pp. 110.
[174] M. Shahabi, S. Member, M. R. Haghifam, S. Member, and
M. Mohamadian, Microgrid dynamic performance improvement using
a doubly fed induction wind generator, IEEE Trans. Energy Convers.,
vol. 24, no. 1, pp. 137145, Mar. 2009.
[175] Wind energy barometer 2011, Systemes SolairesLe Journal de
lolien, vol. 10, no. 8, pp. 80107, Feb. 2011.
[176] Red Elctrica, The Spanish Electricity System. Summary 2008,
last accessed Feb. 2013. [Online]. Available: http://www.ree.es/
ingles/sistema_electrico/informeSEE-2008.asp
[177] Red Elctrica, The Spanish Electricity System. Summary 2009,
last accessed Feb. 2013. [Online]. Available: http://www.ree.es/
ingles/sistema_electrico/informeSEE-2009.asp
[178] Red Elctrica, The Spanish Electricity System. Summary 2010,
last accessed Feb. 2013. [Online]. Available: http://www.ree.es/
ingles/sistema_electrico/informeSEE-2010.asp
[179] Red Elctrica, Press release: Record share of wind energy in demand coverage, last accessed Feb. 2013. [Online]. Available: http://
www.ree.es/ingles/sala_prensa/web/notas_detalle.aspx?id_nota=255
[180] M. Altn, O. Gksu, R. Teodorescu, P. Rodriguez, B.-B. Jensen,
and L. Helle, Overview of recent grid codes for wind power
integration, in Proc. 12th Int. Conf. OPTIM Elect. Electron. Equip.,
2010, pp. 11521160.

CRDENAS et al.: CONTROL SYSTEMS FOR OPERATION OF DFIGs IN WIND ENERGY APPLICATIONS

[181] F. Blaabjerg, R. Teodorescu, M. Liserre, and A. V. Timbus, Overview


of control and grid synchronization for distributed power generation
systems, IEEE Trans. Ind. Electron., vol. 53, no. 5, pp. 13981409,
Oct. 2006.
[182] J. Lopez, P. Sanchis, X. Roboam, and L. Marroyo, Dynamic behavior of the doubly fed induction generator during three-phase voltage
dips, IEEE Trans. Energy Convers., vol. 22, no. 3, pp. 709717,
Sep. 2007.
[183] Red Elctrica, Procedimiento de operacin P.O. 12.3: Requisitos de
respuesta frente a huecos de tensin de las instalaciones de produccin en
rgimen especial, 2006. Last accessed Feb. 2013. [Online]. Available:
http://www.ree.es/operacion/pdf/po/PO_resol_12.3_Respuesta_huecos_
eolica.pdf
[184] Energinet, Technical regulation 3.2.5 for wind power plants with
a power output greater than 11 kW, 2010. Last accessed Feb. 2013.
[Online]. Available: www.energynet.dk; http://www.energinet.dk/EN/El/
Forskrifter/Technical-regulations/Sider/Regulations-for-gridconnection.aspx
[185] I. Erlich, W. Winter, and A. Dittrich, Advanced grid requirements for
the integration of wind turbines into the German transmission system,
in Proc. IEEE Power Eng. Soc. Gen. Meeting, 2006, p. 7.
[186] X. Dawei, R. Li, P. J. Tavner, and S. Yang, Control of a doubly fed induction generator in a wind turbine during grid fault ridethrough, IEEE Trans. Energy Convers., vol. 21, no. 3, pp. 652662,
Sep. 2006.
[187] M. Mohseni, S. M. Islam, and M. A. S. Masoum, Impacts of
symmetrical and asymmetrical voltage sags on DFIG-based wind turbines considering phase-angle jump, voltage recovery, and sag parameters, IEEE Trans. Power Electron., vol. 26, no. 5, pp. 15871598,
May 2011.
[188] S. Hu, X. Lin, Y. Kang, and X. Zou, An improved low-voltage ridethrough control strategy of doubly fed induction generator during grid
faults, IEEE Trans. Power Electron., vol. 26, no. 12, pp. 36533665,
Dec. 2011.
[189] M. Bollen, Understanding Power Quality Problems: Voltage Sags and
Interruptions. Piscataway, NJ, USA: IEEE Press, 1999.
[190] M. Bollen and I. Gu, Signal Processing of Power Quality Disturbances.
Hoboken, NJ, USA: Wiley, 2006.
[191] J. Lopez, E. Gubia, P. Sanchis, X. Roboam, and L. Marroyo, Wind
turbines based on doubly fed induction generator under asymmetrical
voltage dips, IEEE Trans. Energy Convers., vol. 23, no. 1, pp. 321330,
Mar. 2008.
[192] I. Erlich, H. Wrede, and C. Feltes, Dynamic behavior of DFIGbased wind turbines during grid faults, in Proc. Power Convers.
Conf.Nagoya, 2007, pp. 11951200.
[193] J. Morren and S. W. H. de Haan, Short-circuit current of wind turbines
with doubly fed induction generator, IEEE Trans. Energy Convers.,
vol. 22, no. 1, pp. 174180, Mar. 2007.
[194] A. H. Kasem, E. F. El-Saadany, H. H. El-Tamaly, and
M. A. A. Wahab, An improved fault ride-through strategy for
doubly fed induction generator-based wind turbines, IET Renew. Power
Gen., vol. 2, no. 4, pp. 201214, Dec. 2008.
[195] M. Kiani, Effects of voltage unbalance and system harmonics on the
performance of doubly fed induction wind generators, IEEE Trans. Ind.
Appl., vol. 46, no. 2, pp. 562568, Mar./Apr. 2010.
[196] S. Alepuz, A. Calle-Prado, S. Busquets-Monge, S. Kouro, and B. Wu,
Use of stored energy in PMSG rotor inertia for low voltage ride-through
in back-to-back NPC converter based wind power systems, IEEE Trans.
Ind. Electron., vol. 60, no. 5, pp. 17871796, May 2013.
[197] M. K. Das, S. Chowdhury, and S. P. Chowdhury, Protection and voltage
control of DFIG wind turbines during grid faults, in Proc. 10th IET Int.
Conf. DPSP Manag. Change, 2010, pp. 2424.
[198] N. Jouko, Voltage dip ride through of a double-fed generator equipped
with an active crowbar, in Proc. Nordic Wind Power Conf., 2004,
pp. 14091415.
[199] A. Petersson, T. Thiringer, L. Harnefors, and T. Petru, Modeling
and experimental verification of grid interaction of a DFIG wind turbine, IEEE Trans. Energy Convers., vol. 20, no. 4, pp. 878886,
Dec. 2005.
[200] Y. Jun, L. Hui, L. Yong, and C. Zhe, An improved control strategy of
limiting the DC-link voltage fluctuation for a doubly fed induction wind
generator, IEEE Trans. Power Electron., vol. 23, no. 3, pp. 12051213,
May 2008.
[201] J. Hu and Y. He, DFIG wind generation systems operating
with limited converter rating considered under unbalanced network
conditionsanalysis and control design, Renew. Energy, vol. 36, no. 2,
pp. 829847, Feb. 2011.

2797

[202] B. Wu, Y. Lang, N. Zargari, and S. Kouro, Power Conversion and Control
of Wind Energy Systems. Hoboken, NJ, USA: Wiley, 2011.
[203] T. Kawady, C. Feltes, I. Erlich, and A. I. Taalab, Protection system
behavior of DFIG based wind farms for grid-faults with practical considerations, in Proc. IEEE PES Gen. Meeting, 2010, pp. 16.
[204] L. G. Meegahapola, T. Littler, and D. Flynn, Decoupled-DFIG fault
ride-through strategy for enhanced stability performance during grid
faults, IEEE Trans. Sustain. Energy, vol. 1, no. 3, pp. 152132,
Oct. 2010.
[205] C. Feltes, S. Engelhardt, J. Kretschmann, J. Fortmann, F. Koch, and
I. Erlich, High voltage ride-through of DFIG-based wind turbines, in
Proc. 21st IEEE Power Energy Soc. Gen. MeetingConvers. Del. Elect.
Energy, 2008, pp. 18.
[206] K. E. Okedu, S. M. Muyeen, R. Takahashi, and J. Tamura, Participation
of facts in stabilizing DFIG with crowbar during grid fault based on grid
codes, in Proc. IEEE GCC Conf. Exhib., 2011, pp. 365368.
[207] J. A. Suul, M. Molinas, and T. Undeland, STATCOM-based indirect
torque control of induction machines during voltage recovery after grid
faults, IEEE Trans. Power Electron., vol. 25, no. 5, pp. 12401250,
May 2010.
[208] I. Martinez and D. Navarro, Gamesa DAC converter: The way
for REE grid code certification, in Proc. 13th EPE-PEMC, 2008,
pp. 437443.
[209] W. Qiao, G. K. Venayagamoorthy, and R. G. Harley, Real-time implementation of a STATCOM on a wind farm equipped with doubly fed
induction generators, IEEE Trans. Ind. Appl., vol. 45, no. 1, pp. 98
107, Jan./Feb. 2009.
[210] W. Qiao, R. G. Harley, and G. K. Venayagamoorthy, Coordinated reactive power control of a large wind farm and a STATCOM using heuristic
dynamic programming, IEEE Trans. Energy Convers., vol. 24, no. 2,
pp. 493503, Jun. 2009.
[211] C. Wessels, F. Gebhardt, and F. W. Fuchs, Fault ride-through of a DFIG
wind turbine using a dynamic voltage restorer during symmetrical and
asymmetrical grid faults, IEEE Trans. Power Electron., vol. 26, no. 3,
pp. 807815, Mar. 2011.
[212] A. Teninge, D. Roye, and S. Bacha, Reactive power control for variable
speed wind turbines to low voltage ride through grid code compliance,
in Proc. XIX ICEM, 2010, pp. 16.
[213] G. Joos, Wind turbine generator low voltage ride through requirements and solutions, in Proc. 21st IEEE Power Energy Soc. Gen.
MeetingConvers. Del. Electrical, 2008, pp. 17.
[214] G. Pannell, D. J. Atkinson, and B. Zahawi, Minimum-threshold
crowbar for a fault-ride-through grid-code-compliant DFIG wind
turbine, IEEE Trans. Energy Convers., vol. 25, no. 3, pp. 750759,
Sep. 2010.
[215] W. Guo, L. Xiao, and S. Dai, Enhancing low-voltage ride-through
capability and smoothing output power of DFIG with a superconducting fault-current limitermagnetic energy storage system, IEEE Trans.
Energy Convers., vol. 27, no. 2, pp. 277295, Jun. 2012.
[216] A. Dittrich and A. Stoev, Comparison of fault ride-through strategies
for wind turbines with DFIM generators, in Proc. Eur. Conf. Power
Electron. Appl., 2005, p. 8.
[217] S. Heier, Grid Integration of Wind Energy Conversion Systems.
Hoboken, NJ, USA: Wiley, 2006.
[218] O. Gomis-Bellmunt, A. Junyent-Ferr, A. Sumper, and J. Bergas-Jan,
Ride-through control of a doubly fed induction generator under unbalanced voltage sags, IEEE Trans. Energy Convers., vol. 23, no. 4,
pp. 10361045, Dec. 2008.
[219] L. Trilla, O. Gomis-Bellmunt, A. Junyent-Ferre, M. Mata, J. Sanchez
Navarro, and A. Sudria-Andreu, Modeling and validation of DFIG
3-MW wind turbine using field test data of balanced and unbalanced
voltage sags, IEEE Trans. Sustain. Energy, vol. 2, no. 4, pp. 509519,
Oct. 2011.
[220] S. Engelhardt, J. Kretschmann, J. Fortmann, F. Shewarega, I. Erlich, and
C. Feltes, Negative sequence control of DFG based wind turbines, in
Proc. IEEE Power Energy Soc. Gen., 2011, pp. 18.
[221] J. Lopez, P. Sanchis, E. Gubia, A. Ursua, L. Marroyo, and X. Roboam,
Control of doubly fed induction generator under symmetrical voltage
dips, in Proc. IEEE ISIE, 2008, pp. 24562462.
[222] J. Liang, W. Qiao, and R. G. Harley, Feed-forward transient current
control for low-voltage ride-through enhancement of DFIG wind turbines, IEEE Trans. Energy Convers., vol. 25, no. 3, pp. 836843,
Sep. 2010.
[223] J. Lopez, E. Gubia, E. Olea, J. Ruiz, and L. Marroyo, Ride through
of wind turbines with doubly fed induction generator under symmetrical
voltage dips, IEEE Trans. Ind. Electron., vol. 56, no. 10, pp. 42464254,
Oct. 2009.

2798

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 7, JULY 2013

[224] Y. Wang, L. Xu, and B. W. Williams, Improved operation of DFIG


and FSIG-based wind farms during network unbalance, in Proc. 21st
IEEE Power Energy Soc. Gen. MeetingConvers. Del. Elect., 2008,
pp. 17.
[225] J. Hu, H. Nian, H. Xu, and Y. He, Dynamic modeling and improved
control of DFIG under distorted grid voltage conditions, IEEE Trans.
Energy Convers., vol. 26, no. 1, pp. 163175, Mar. 2011.
[226] M. Mohseni, S. Islam, and M. A. S. Masoum, Fault ride-through capability enhancement of doubly-fed induction wind generators, IET
Renew. Power Gen., vol. 5, no. 5, pp. 368376, Sep. 2011.
[227] M. Mohseni and S. Islam, Transient control of DFIG-based wind power
plants in compliance with the australian grid code, IEEE Trans. Power
Electron., vol. 27, no. 6, pp. 28132824, Jun. 2011.
[228] M. I. Martinez, G. Tapia, A. Susperregui, and H. Camblong, Slidingmode control for DFIG rotor- and grid-side converters under unbalanced
and harmonically distorted grid voltage, IEEE Trans. Energy Convers.,
vol. 27, no. 2, pp. 328339, Jun. 2012.

Roberto Crdenas (S95M97SM07) was born


in Punta Arenas, Chile. He received the B.S. degree
from the University of Magallanes, Punta Arenas, in
1988 and the M.Sc. and Ph.D. degrees from the University of Nottingham, Nottingham, U.K., in 1992
and 1996, respectively.
From 1989 to 1991 and from 1996 to 2008, he
was a Lecturer with the University of Magallanes.
From 1991 to 1996, he was with the Power Electronics Machines and Control Group, University of
Nottingham. From 2009 to 2011, he was with the
Department of Electrical Engineering, University of Santiago, Santiago, Chile.
He is currently a Professor of power electronics and drives with the Department
of Electrical Engineering, University of Chile, Santiago. His main research
interests include the control of electrical machines, variable-speed drives, and
renewable energy systems.
Dr. Crdenas was a recipient of the Best Paper Award from the IEEE
T RANSACTIONS ON I NDUSTRIAL E LECTRONICS in 2005 and the Ramon
Salas Edward Award from the Chilean Institute of Engineers in 2009.

Rubn Pea (S95M97) was born in Coronel,


Chile. He received the Electrical Engineering degree
from the University of Concepcion, Concepcion,
Chile, in 1984 and the M.Sc. and Ph.D. degrees from
the University of Nottingham, Nottingham, U.K., in
1992 and 1996, respectively.
From 1985 to 2008, he was a Lecturer with
the University of Magallanes, Punta Arenas, Chile.
He is currently with the Department of Electrical
Engineering, University of Concepcin. His main
research interests include the control of power electronics converters, ac drives, and renewable energy systems.

Salvador Alepuz (S98M03SM12) was born


in Barcelona, Spain. He received the M.Sc. and
Ph.D. degrees in electrical and electronic engineering from the Technical University of Catalonia (UPC), Barcelona, Spain, in 1993 and 2004,
respectively.
Since 1994, he has been an Associate Professor
with the Matar School of Technology (Tecnocampus Matar-Maresme), UPC, Matar, Spain. From
2006 to 2007, he was with the Departamento de Electrnica, Universidad Tcnica Federico Santa Mara,
Valparaso, Chile, conducting postdoctoral research. In 2009, he was a Visiting
Researcher for three months with the Department of Electrical and Computer
Engineering, Ryerson University, Toronto, ON, Canada. His research interests
include multilevel conversion and ac power conversion applied to renewable
energy systems.

Greg Asher (M98SM04F07) received the B.Sc.


and Ph.D. degrees from Bath University, Bath, U.K.,
in 1976.
He is a Research Fellow in superconducting systems with the University of Bangor, Gwynedd, U.K.
In 1984, he was appointed as a Lecturer of control with the University of Nottingham, Nottingham,
U.K. In 2000, he was appointed as a Professor of
electrical drives; in 2004, as a School Head with the
School of Electrical and Electronic Engineering; and
in 2008, as the Associate Dean for Teaching and
Learning with the Faculty of Engineering, University of Nottingham. He is
the author of nearly 300 research papers. He has received over 5 million
in research contracts and has successfully supervised 31 Ph.D. students. His
research interests include motor drive control, cover power system modeling,
power microgrid control, aircraft power systems, and motor drive systems,
particularly the control of ac machines.
Dr. Asher was a member of the Executive Committee of the European Power
Electronics Association until 2003 and the Chair of the Power Electronics
Technical Committee of the IEEE Industrial Electronics Society until 2008.
He is currently an Associate Editor for the IEEE Industrial Electronics Society.

You might also like