Download as pdf or txt
Download as pdf or txt
You are on page 1of 38

The current issue and full text archive of this journal is available at

www.emeraldinsight.com/0332-1649.htm

COMPEL SPECIAL ISSUE PAPER


32,1
Electrical machines and
power-electronic systems for
34
high-power wind energy
generation applications
Part II – power electronics and control systems
Z.Q. Zhu and Jiabing Hu
Department of Electronic and Electrical Engineering,
The University of Sheffield, Sheffield, UK
Abstract
Purpose – Power-electronic systems have been playing a significant role in the integration of
large-scale wind turbines into power systems due to the fact that during the past three decades
power-electronic technology has experienced a dramatic evolution. This second part of the paper aims
to focus on a comprehensive survey of power converters and their associated control systems for
high-power wind energy generation applications.
Design/methodology/approach – Advanced control strategies, i.e. field-oriented vector control
and direct power control, are initially reviewed for wind-turbine driven doubly fed induction generator
(DFIG) systems. Various topologies of power converters, comprising back-to-back (BTB) connected
two- and multi-level voltage source converters (VSCs), BTB current source converters (CSCs) and
matrix converters, are identified for high-power wind-turbine driven PMSG systems, with their
respective features and challenges outlined. Finally, several control issues, viz., basic control targets,
active damping control and sensorless control schemes, are elaborated for the machine- and grid-side
converters of PMSG wind generation systems.
Findings – For high-power PMSG-based wind turbines ranging from 3 MW to 5 MW,
parallel-connected 2-level LV BTB VSCs are the most cost-effective converter topology with mature
commercial products, particularly for dual 3-phase stator-winding PMSG generation systems. For
higher-capacity wind-turbine driven PMSGs rated from 5 MW to 10 MW, medium voltage multi-level
converters, such as 5-level regenerative CHB, 3- and 4-level FC BTB VSC, and 3-level BTB VSC, are
preferred. Among them, 3-level BTB NPC topology is the favorite with well-proven technology and
industrial applications, which can also be extensively applicable with open-end winding and dual
stator-winding PMSGs so as to create even higher voltage/power wind generation systems. Sensorless
control algorithms based on fundamental voltages/currents are suggested to be employed in the basic
VC/DPC schemes for enhancing the robustness in the entire PMSG-based wind power generation
system, due to that the problems related with electromagnetic interferences in the position signals and
the failures in the mechanical encoders can be avoided.
Originality/value – This second part of the paper for the first time systematically reviews the latest
state of arts with regard to power converters and their associated advanced control strategies for
high-power wind energy generation applications. It summarizes a variety of converter topologies with
COMPEL: The International Journal
for Computation and Mathematics in pros and cons highlighted for different power ratings of wind turbines.
Electrical and Electronic Engineering
Vol. 32 No. 1, 2013
Keywords Control, Doubly fed induction generator (DFIG),
pp. 34-71 Permanent magnet Synchronous Generator (PMSG), Power converters, Wind energy, Wind power,
q Emerald Group Publishing Limited Electric power generation
0332-1649
DOI 10.1108/03321641311293740 Paper type General review
1. Introduction High-power
As the massive penetrations of wind turbines worldwide, wind power technology has wind energy
also undergone a dramatic transformation in the last decade, with proven technology
for turbines, electrical machines as well as power-electronic systems. Variable-pitch generation
control and variable-speed operation are well developed for turbines and generators,
respectively, to improve the efficiency and reliability of the power generation systems.
As a consequence, a large number of variable-speed pitch-controlled wind turbines are 35
installed in modern wind farms. At present, most installed wind turbine generators are
in the power range of 1.5-5 MW, and the gear-drive doubly-fed induction generator
(DFIG)-based wind turbines are still dominating the markets for the time being. On the
other hand, numerous research efforts have been made for larger systems, targeting
5-10 MW level. In these wind power generation systems, a direct-drive permanent
magnet synchronous generator (PMSG) supplied with a full-capacity power converter
has been proven to be a more effective solution compared with the DFIG-based system,
in particular for off-shore wind farms, due to its low maintenance cost without gear
boxes, complete decoupling from the grid, wider speed range and enhanced capability
of fault-ride-through operations.
Meanwhile, power converters are necessitated in modern wind power generation
systems for the purpose of variable-speed operation, independent control of
active/reactive powers and easy maximum-power-point-tracking (MPPT) operations.
In addition, the increasing penetration of wind turbines requires new control strategies
so as to maintain or even to enhance the quality and stability of the connected electrical
power systems, which have been defined as modern grid codes depicted in the first
part, for large-scale wind farms.
Power-electronic systems have been playing an important role in the integration of
large-scale wind turbines into the grid due to the fact that power-electronic technology
has experienced a rapid evolution, mainly in three factors, during the past three decades.
The first one is the advent of self-commutated semiconductor switches (IGBT, integrated
gate-commutated thyristor (IGCT), SiC FET, etc.) with high-frequency switching, which
are capable of carrying high breakdown voltage/current and as a result handling high
powers. The second factor is the development of advanced control strategies, such as
field-oriented control (FOC) (Pena et al., 1996; Petersson et al., 2005), direct torque/power
control (DTC/DPC) (Xu and Cartwright, 2006; Abad et al., 2008a), predictive control
(Abad et al., 2008a, b), nonlinear control (Beltran et al., 2008; Hu et al., 2010), etc. associated
with diversity of PWM techniques to improve the system responses and power qualities.
The last one is the introduction of high-speed real-time digital signal processors (DSPs)
and microcomputers that can implement advanced and complex control algorithms.
These three factors together have successfully led to the advent of a massive variety of
cost-effective and grid-friendly converters for collecting wind power into the grid.
Recently, a few reviews (Baroudi et al., 2005; Carrasco et al., 2006; Chen et al., 2009)
were reported on various types of conceptual power converter topologies for wind
generators, including:
.
out-of-date diode rectifier and thyristor grid-side inverter for PMSG (Chen and
Spooner, 1998, 2003);
.
diode rectifier, dc/dc boost converter and voltage source inverter for PMSG
(Higuchi et al., 2000; Song et al., 2003; Haque et al., 2008a, b, 2010) and EESG
(Enercon GmbH, 2009);
COMPEL .
classic two-level back-to-back (BTB) connected voltage source converters (VSCs)
32,1 for DFIG (Pena et al., 1996; Muller et al., 2002), SCIG (Bueno et al., 2008) and
PMSG (Chinchilla et al., 2006a, b; Ramtharan et al., 2007); and
.
matrix converter for DFIG (Zhang et al., 1997).

However, control strategies have been rarely surveyed on power converters for either
36 DFIG or PMSG-based wind power generation systems. As one of the main aims of this
paper, this second part attempts to comprehensively identify several well-proven and
potential power-electronic systems with their advanced control strategies for
high-power (3-10 MW) wind energy applications. The rest of this paper is organized
as follows. In Section 2, advanced power-electronic control strategies, including classic
vector control (VC), conventional look-up-table direct power control (LUT-DPC) and
potential sliding mode control (SMC)-based DPC, are briefly reviewed, for DFIG-used
grid- and rotor-side converters (RSCs), respectively. In Section 3, various topologies of
power converters for high-power PMSG systems are identified, including a variety of
parallel-connections of two-level low-voltage (LV) BTB VSCs, different medium-voltage
(MV) multi-level VSCs (five-level regenerative cascaded H-bridge (CHB),
three-/four-level BTB flying capacitor (FC) and three-level BTB neutral-point-clamped
(NPC)), BTB-connected current-source converters (CSCs) and matrix converters. As to
the control systems for the PMSG-used machine- and grid-side converters (GSCs), classic
control targets, active damping control strategies and two sensorless control schemes
are highlighted in Section 4. Finally, Section 5 draws some conclusions.

2. Power-electronic control systems for wind-turbine driven DFIG


Generally, the DFIG-used converter employs two three-phase two-level VSCs (denoted
as GSC and RSC) with forced-commutated switches (i.e. IGBTs or IGCTs), which are
connected BTB via a common dc-link capacitor (Muller et al., 2002), as shown in Figure 1.
By way of example, for a 5 MW DFIG system, its converter is rated at only 1 MW with
speed range varying from 0.8 to 1.2 pu, as a result, two-level LV VSC with currently
available semiconductor devices is adequate to handle such generator operation.
Owing to the decoupled capability of the dc-link capacitor, the GSC and RSC play
their roles independently. The main role of the GSC is to maintain the dc-link voltage
constant, to control the power factor of GSC (usually at unity power factor under normal

Rotor-side Grid-side
converter (RSC) converter (GSC)

Line
inductors Transformer Ac grid
(necessary)
Gear
DFIG
box

Figure 1.
Diagram of a wind-turbine
driven DFIG based on a
two-level BTB VSC
grid-voltage conditions) and to keep the ac current sinusoidal with reduced harmonic High-power
contents injected into the grid. While for the RSC, it is responsible for regulating the DFIG’s wind energy
stator active and reactive power outputs without coupling, accomplishing the variable-speed
constant-frequency operation and obtaining MPPT during wind speed changes. generation
In order to obtain high-quality system performance, as in ac motor drives, VC and
DTC/DPC strategies have also widely been employed in the field of controlling
wind-turbine driven DFIGs. In terms of VC with different orientations, it has 37
voltage-oriented and flux-oriented schemes. For instance, GSC can be controlled either
by grid-voltage-oriented (GVO) VC scheme or via virtual-flux-oriented (VFO) VC
scheme, while for RSC, stator-flux-oriented (SFO) and stator-voltage-oriented (SVO) VC
schemes can be employed, respectively. As to direct power/torque control for GSC and
RSC, it can be divided into look-up-table (LUT) method and SVM based means, which
provide variable and constant switching frequencies, respectively. For clear
illustrations, Table I summarizes various VC and DPC/DTC strategies for
DFIG-used GSC and RSC, respectively, reported in the existing literature.
Taking SVO-based VC strategy (He et al., 2005) for RSC as an example, the control
diagram is shown in Figure 2. As shown, with stator-voltage orientation, the scheme

GSC RSC

VC GVO (Choi and Sul, 1998) SVO (Petersson, 2005; He et al., 2005)
VFO (Malinowski, 2001) SFO (Pena et al., 1996; He et al., 2005)
DPC LUT (Larrinaga et al., 2007; LUT (Xu and Cartwright, 2006; Datta Table I.
Vazquez et al., 2008) and Ranganathan, 2001) Different VC and DPC
Space-vector-modulation (SVM) Space-vector-modulation (SVM) control strategies for
(Malinowski et al., 2004) (Zhi and Xu, 2007) DFIG’s GSC and RSC

Grid side converter Rotor side converter


Sabc
SVM

ej(q1-q1)

DFIG Decoupling
Isabc Irabc

qr wr 3s/2s
Irdq
d/dt Ird, Irq
encoder e–j(q1-qr) PI controller
Issab yssab
Grid Stator flux ysdq
3s/2s
estimation Issab e–jq1
q1 Isdq
PLL I*rdq
Usdq
Usabc s Ps, Qs Figure 2.
3s/2s U sab Ps, Qs
Ps, Qs SVO-based VC scheme
calculation P*s, Q*s for RSC of DFIG
PI controller
generation system
Source: He et al. (2005)
COMPEL decouples the ac rotor current into active and reactive power components (Ird and Irq) in
the synchronous reference frame rotating at the angular speed of grid voltage. Control
32,1 of instantaneous stator active and reactive powers is then achieved by regulating the
decoupled rotor currents, respectively, using proportional-integral (PI) controllers. The
reference values of d- and q-axis rotor currents are obtained, respectively, from
outer-loop PI controllers for stator active and reactive powers.
38 According to the principle of the LUT DPC for a grid-connected DFIG (Xu and
Cartwright, 2006), at each sampling instant an appropriate voltage vector is selected by
the switching rule to restrict the instantaneous active and reactive powers within their
required hysteresis bands, respectively. The scheme is shown in Figure 3. The active
and reactive power controllers are three-level hysteresis comparators, as shown in
Figure 4. The comparators generate discrete signals, Sp and Sq, as inputs to the
switching table based on the active and reactive power errors. With the error signs
from the hysteresis controllers and referring to the sector where the estimated stator
flux is located, the switching rule directly produces the converter’s switching signals
Sabc from a pre-defined LUT. The DPC algorithm is accomplished in the rotor reference

Grid side converter Rotor side converter

Sabc

Optimal
DFIG switching
table
Usabc Isabc
qr wr
d/dt
encoder
Issab y ssab y rsab
Grid Stator flux
3s/2s Stator flux
estimation e–jqr position
Ps, Qs Ps, Qs
s Ps, Qs
Figure 3. 3s/2s U sab hysteresis
calculation P*s, Q*s comparator Sp, Sq
Diagram of traditional
LUT-DPC for DFIG’s RSC
Source: Xu and Cartwright (2006)

Sp Sq

1 1
Pserror –DQ Qserror
–DP
0 DP 0 DQ

Figure 4. –1 –1
Active/reactive power
hysteresis comparators
Source: Xu and Cartwright (2006)
frame rotating at the angular speed of vr without involving any PWM module, which High-power
gets the maximum dynamic performance available. wind energy
However, like a basic DTC, LUT-based DPC has switching frequencies varying
significantly with active/reactive power variations, the power controllers’ hysteresis generation
bands as well as the machine operating velocity. As a result, the stator side ac filter
preventing switching harmonics from injecting the connected grid needs to be
designed to absorb broadband harmonics and the filter’s efficiency is reduced with 39
increased size and power losses. To solve this issue highlighted, in Abad et al., 2008a, b,
the switching vectors were chosen based on a basic switching table and thereafter their
duration times were optimized with the target of reducing pulsations in the
torque/active power and flux/reactive power. Although a constant switching frequency
was achieved, it required complicated online calculations and had oscillating problems
when the generator operates around its synchronous speed. A simple constant
switching frequency DPC strategy based on a predictive power model was developed
in Zhi and Xu (2007) and Hu et al. (2008a). The method, however, was implemented in
the synchronous reference frame, which necessitates the angular information of
network voltage and the synchronous coordinate transformations.
More recently, the authors proposed a new DPC, respectively, for GSC (Hu et al.,
2011) and RSC (Hu et al., 2010) of grid-connected DFIG systems based on SMC
approach. Taking SMC-DPC for DFIG’s RSC as an example, the schematic diagram is
shown in Figure 5. As shown, the developed SMC directly generates the voltage
reference for RSC in the stator stationary reference frame according to the
instantaneous errors of active and reactive powers. Afterwards, it is transformed into
rotor reference frame, and SVM module is used to generate the required switching
voltage vectors and their respective duration times. Compared with classic VC and
conventional LUT-DPC strategies, SMC-DPC features (Hu et al., 2010):
.
no synchronous coordinate transformations and angular information of grid
voltage or stator flux are required;
.
enhanced transient performance similar to the conventional LUT-DPC is obtained;

Grid side converter Rotor side converter


Srabc
SVPWM

DC link
Fitter r
capacitor Urab

(αβ)s/(αβ)r
DFIG
encoder qr wr

Irabc
Usabc Isabc U srab
I rrab I srab
abc/αβ (αβ)r/(αβ)s
Network Ussab I ssab Figure 5.
abc/αβ DPC algorithm based on sliding Diagram of SMC-DPC
P*s, Q*s mod control approach for RSC of a wind
turbine DFIG
Source: Hu et al. (2010)
COMPEL .
steady-state stator and rotor current harmonic spectra are kept at the same level
32,1 as the classic VC strategy due to the use of SVM modulation technique; and
.
the steady-state and transient responses are insensitive to the machine
parameters’ variations.

It can be concluded that the developed SMC-DPC is pretty promising in the application
40 of wind-turbine driven DFIG’s GSC and RSC.
On the other hand, as pointed out in Table II of part I, since the DFIG’s stator is
directly connected to grid via a step-up transformer, the generation system is very
susceptible to grid disturbance, such as voltage dip, phase-angle jump, three-phase
unbalances and harmonically distortions. There exit significant challenges to the
power-electronic control systems for wind-turbine driven DFIGs, in order to ride
through these grid faults, namely, stay connected to the grid and contribute to
the recovery of faulted grid. Among them, LV ride through capability, is the most
challenging and has obtained massive interests in research and development by
academic researchers (He et al., 2005; Morren and de Haan, 2005; Hansen et al., 2006)
and industrial companies (GE, 2010; Saylors, 2008). Meanwhile, asymmetric grid faults
occur more frequently than symmetric ones. If a DFIG control system does not take
into account the voltage unbalance, the stator and rotor currents could be highly
unbalanced and distorted even under a slight stator-voltage unbalance. It was reported
in Muljadi et al. (1999) that 30 percent negative sequence current was caused by
6 percent negative sequence voltage in a prototype induction generator. The
unbalanced currents create unequal heating in the three-phase stator/rotor windings
and torque pulsations on the generator system. As a result, the DFIG-based wind
turbines without unbalanced voltage control might have to be disconnected from the
grid during network voltage unbalance (Nass et al., 2002). It has been found that wind
farms connected to distribution networks periodically experience larger than 2 percent
voltage unbalance, which has caused a large number of trips (Codd, 2003). On the other
hand, the emerging grid codes require that the wind farm connected to a distribution
network should be able to withstand a maximum value of 2 percent steady-state
voltage unbalance and relatively large transient voltage unbalance without tripping
(National Grid Transco, 2004). System control and operation of the wind-turbine driven
DFIG under unbalanced grid-voltage conditions have been thoroughly investigated in
Xu and Wang (2007), Hu et al. (2009) and Hu and He (2009). Consequently, with the
coordinated control strategy proposed in Hu and He (2009) the performance and
operation of DFIG-based wind power generation system under unbalanced conditions
can be significantly improved by simultaneously eliminating the oscillations in the
electromagnetic torque and the total active or reactive power, or the current imbalance
generated from the overall system.

3. Power-electronic systems for high-power direct-drive PMSG


As given in Table I of part I, although DFIG-based wind turbines are still dominating
the market for the time being, full-power-converter-supplied EESGs and especially
PMSGs have been favored by many turbine manufacturers for high-power, direct-drive
and off-shore wind farm applications. It was presented in Khan et al. (2009) that with
uncontrolled diode rectifier as machine-side converter of a PMSG system, the
generator’s voltage and current are considerably distorted, and the current total
harmonic distortion (THD) keeps high throughout the whole rotor speed, which leads High-power
to quite low efficiency with significant harmonic losses. As a result, it can be concluded wind energy
that the diode rectifier is quite fit for cost-effective small scale PMSG-based wind
power system, but not suitable for MW-ranking wind-turbine driven PMSG (especially generation
direct-drive applications) due to its considerable current distortions, periodic torque
oscillations and harmonic losses.
Thus, in this section, various topologies of power converters with full-controlled 41
rectifiers as machine-side converters for PMSG systems are identified and reviewed.
A variety of parallel-connections of two-level LV BTB VSCs, different MV multi-level
VSCs (five-level regenerative CHB, three-/four-level BTB FC and three-level BTB NPC),
BTB-connected CSCs and matrix converters are included, which are classified in
Figure 6 for quick reference.

3.1 Parallel-connected LV BTB VSC topologies


More recently, 3-5 MW PMSG-based wind turbines have been under test/construction
mainly for off-shore wind farms. As a result, the power rating of converters must be
even larger than that of the generators for safety purpose. Such high-rating converters
can be realized either by parallel-connected three-phase two-level BTB VSCs with LV
and high current or via multi-level VSCs with MV and low current. At present, both
generators and converters are usually designed at LVs (, 1 kV, usually 690 V
line-to-line voltage) due to economical and technical reasons. While in the near future,
once the MV generators are manufactured, multi-level converters will be favored,
which is addressed in the next sub-section.
This sub-section is to briefly review the possible topologies on parallel connection of
three-phase two-level LV BTB VSCs for wind-turbine driven PMSGs. Taking
two-parallel connection as an example, Figure 7(a)-(e) (Li et al., 2008) shows the various
combinations, respectively.
As shown in Figure 7(a), two identical BTB VSCs are parallel-connected between
the PMSG and the grid with individual (isolated) dc-link capacitors, as a result,

PMSG Transformer Grid


Gear box Full power
(optional) converter

High-Power (3MW ~ 10MW) PMSG Converters

Low-Voltage (LV) Medium Voltage (MV)


(<1kV) Voltage-Source (2.3kV ~ 13.8kV) Converters
Converters (VSCs)

BTB-connected PWM MV VSCs based on LV-


MV Current-Source IGBTs or MV-IGCTs/IGBTs
Parallel-Connection
Converters (CSCs)
2-Level back-to-back
(BTB) VSCs based Multilevel VSCs 2-level VSCs
on LV IGBTs Figure 6.
5-Level MV/HV Series- Potential high-power
BTB-Connected BTB-Connected
Cascaded H- 3/4-Level Flying IGBTs Connected converter topologies for
3-Level Neutral- LV IGBTs
Bridge VSCs Capacitor (FC) Point-Clamped full-power-supplied PMSG
VSCs (NPC) VSCs wind generation systems
COMPEL MSC 1 Two dc-link
capacitor
GSC 1 Converter 1
3 phase
MSC 1
ONE dc-link
capacitor GSC 1

32,1 3 phase

PMSG
Ls
C
Lg Transformer Ac grid
PMSG
C
Lg Transformer

Ls Ac grid

is0 ig0

Ls Lg Ls1 Lg1
C Current
Current

42 sharing sharing
reactor reactor

MSC 2 GSC 2 Converter 2 MSC 2 GSC 2

(a) With individual dc capacitor (Li et al. (2008)) (b) With ONE dclink and current sharing reactors
(according to (Hu et al. (2008b)))

MSC 1 ONE dc-link GSC 1 Six phase


capacitor
One ONE dc-link
half-bridge MSC capacitor GSC 1
module
3 phase
Lg Ac grid 6 phase
Ls Transformer Transformer Ac grid
Ls Lg
Figure 7. PMSG
C
PMSG
C

Various parallel is0 ig0


ig0

connections of three-phase
two-level LV BTB VSCs Ls Lg Ls Lg
C
for high-power
wind-turbine driven
PMSGs (two parallel as MSC 1 GSC 2 GSC 2
an example)
(c) With ONE dc link capacitor (d) 6-phase PMSG withONE dc link

the zero-sequence circulating current problem does not exist. Carrier phase shifted
(CPS) (Li et al., 2008) or interleaved (Wen et al., 2009; Tang et al., 2009) modulation
techniques can be used for both MSCs and GSCs to reduce current ripples/harmonics
under low switching frequencies (1-2 kHz). N þ 1 redundant operation is possible but
with entire BTB VSC module, which, together with the isolated dc-link capacitors,
leads to high cost and large bulk of entire parallel system.
According to Matsui et al. (1993), a two-parallel connection of three-phase two-level
VSCs can be constructed in Figure 7(b) for a wind-turbine driven PMSG. As shown,
only one dc-link is required and two respective GSCs and MSCs are parallel-connected
through current-sharing reactors, which are employed to balance current, but a
zero-sequence circulating current occurs due to practical asymmetry of reactors’ value.
As a result, a synchronized control was proposed in Ogasawara et al. (1992), which
basically treats the parallel converters as one converter (six phase) to nullify
circulating current. For example, two-parallel three-phase three-leg converters are
controlled as a three-phase six-leg converter. This method is not suitable for modular
converter design. Besides, when more converters are in parallel, the system becomes
very complicated to design and control.
Figure 7(c) shows a non-isolated parallel connection with one dc-link where
individual converters connect both ac and dc sides directly without additional passive
components (Hu et al., 2008b). As shown, four identical converters with their own
filtering inductances are used as two MSCs and GSCs, respectively. Meanwhile,
control strategies presented in Section 4 can be designed within each corresponding
converter so as to facilitate modular design. Further, N þ 1 operation is available
with two VSC modules that are parallel-connected with dc-link, respectively, for MSCs
or GSCs (Qiu and Chen, 2007). CPS or interleaved modulation techniques can be
used to reduce current ripples, and additional controller for nullifying zero-sequence
circulating current is required in the control system (Wen et al., 2009; Tang et al., 2009; High-power
Ye et al., 2002). wind energy
When PMSG is designed with multi-phase (typically multi three phases), such as six
or nine phases (Vizireanu et al., 2006; Kato et al., 2006), a multi-phase two-level generation
half-bridge VSC can be employed as MSC, as shown in Figure 7(d) (six phase as an
example). As a result, control strategies and modulation techniques on multi-phase
drive systems (Zheng et al., 2008) can be transplanted to the multi-phase wind-turbine 43
driven PMSGs. Meanwhile, GSC also utilizes the same direct parallel connection as that
shown in Figure 7(c). Alternatively, two isolated dc-links can be employed, which is
shown in Figure 14(b) of part I.
For clear purpose, the main features of the parallel-connected two-level LV BTB
VSCs, especially as shown in Figures 7(c) and (d) and 14(b) of part I, for high-power
wind-turbine-direct-drive PMSG systems can be summarized as follows:
.
maturely modular design and low dv/dt;
.
high reliability and flexible N þ 1 redundancy;
.
low switching frequency with interleaved modulations and low switching losses;
and
.
low manufacturing expense due to LV power devices.

On the contrary, the main drawbacks of the high-power PMSG-used parallel-connected


LV VSCs are also apparent, namely:
. circulating currents are produced with interleaved modulation techniques; and
.
step-up transformers are necessitated.

3.2 Multi-level MV BTB VSC topologies


Higher capacity of the wind turbine (5-10 MW) necessitates the operation of
corresponding generators and power converters at higher voltages (MV range,
2.3-13.8 kV). Operation of the two-level converter system shown in Figure 7 at high
voltage levels requires device valves with high voltage ratings. This can be achieved
either by using the state-of-the-art power devices at high voltage ratings (3.3, 4.5 and
6.5 kV IGBT modules), or by series connection of LV switches. The former option
imposes excessive switch loss which may render the converter unit economically
unattractive or even unacceptable. Furthermore, as indicated in Krug et al. (2007), the
two-level VSC on the basis of the 6.5 kV IGBT modules cannot be extended in
industrial applications where a high converter efficiency (e.g. h ¼ 99 percent) and a
low THD of the output voltage (e.g. THD # 5 percent) are required since an LC sine
filter cannot be realized at low carrier frequencies (e.g. 450 Hz). Furthermore, the
two-level VSC is not attractive for high switching frequency MV applications since the
high switching losses of the 6.5 kV IGBT modules significantly limit the switch
utilization and the maximum witching frequency.
The latter as shown in Figure 8, where a BTB two-level high-power VSC
employs three series-connected devices in each half leg, exposes its own technical
challenges, i.e.:
.
equal voltage sharing of each series device;
.
strictly simultaneous gating requirements;
COMPEL MSC GSC
32,1
Vdc
Transformer
Ls Lg (optional)
Ac grid
44 PMSG
C

Figure 8.
Wind-turbine driven
PMSG based on two-level
high-power BTB VSCs

.
considerably high dv/dt and high wave reflection;
.
high generator-side filter inductances required to reduce harmonics to
generators;
.
high grid filter inductances required to fulfill harmonic standards by grid codes;
and
.
LCL filters can be used, but introduce possible resonances.

As a result, another approach to avoid the above problems is the use of multi-level VSC
instead of two-level one. Highly popular are the multi-level VSCs, which can be divided
into three categories, according to their topology: NPC (Nabae et al., 1981), FC
(Meynard and Foch, 1992) and CHB (Marchesoni et al., 1988).
An overview of available pulse-width-modulated industrial MV drives on the
market is provided in the Appendix in part I.
In the following sub-sections, potential full-capacity BTB-connected VSC schemes
for wind-turbine driven PMSG systems with the above-mentioned three multi-level
topologies and their variants are identified.
A. Five-level CHB as MSC. The CHB multi-level inverter appeared first in 1988
(Marchesoni et al., 1988), matured during 1990s and gained more attention after 1997.
Currently, this topology has been used in MV high-power drives (maximum rates of
13.8 kV, 1,400 A and 31 MVA) due to its modular structure and power-quality
operational characteristics. It is worth pointing out that in the MV drives without
regenerative operation, three-phase-diode half-bridge rectifiers are used to provide
isolated dc voltage source in each power cell and multi-pulse transformers with
appropriate secondary winding displacements are utilized to reduce input-current
harmonics. The CHB inverter with seven or nine voltage levels has been increasingly
used in high-power MV drives, where the IGBTs are exclusively used as switching
devices. However, no reports are found yet on its application in the field of wind power,
particularly direct-drive PMSG systems.
Figure 9 shows two potential CHB schemes for PMSG’s full-power converters,
where five-level CHB is used as machine-side converter, and three- or single-phase
two-level VSCs can be employed as grid-connected converter in each power cell,
respectively.
The CHB multi-level converter has a number of features and drawbacks, including: High-power
.
Modular structure. The multi-level converter is composed of multiple units of wind energy
identical H-bridge power cells, which leads to a reduction in manufacturing cost. generation
.
Lower voltage THD and dv/dt. The CHB output voltage waveform is formed by
several voltage levels with small voltage steps. Compared with a two-level
converter in Figure 8, the CHB multi-level inverter can produce an output voltage
with much lower THD and dv/dt.
45
.
High-voltage operation without switching devices in series. The H-bridge power cells
are connected in cascade to produce high ac voltages. The problems of equal voltage
sharing for series-connected devices, as shown in Figure 8, are eliminated.
.
Flexible redundant converter design. Actually, the CHB VSC is the only
commercially available converter that offers redundancy as an option. In
this case, one additional power cell per phase and a mechanical bypass per cell,
which short-circuits a failed power cell in 250 ms, are added to the converter
structure.
.
Large number of isolated dc supplies. As shown in Figure 9, in each power cell,
the dc supplies for the CHB converter are usually obtained from a single-phase
full-bridge VSC or three-phase half-bridge VSC.
.
High component count. The CHB converter uses a large number of IGBT
modules, especially in PMSG application as shown in Figure 9. A five-level CHB
converter with single-/three-phase VSC as dc supplier requires 48/60 IGBTs with
the same number of gate drivers.

B. Three/four-level BTB FC-VSC for PMSG. The FC-VSC topology was proposed about
18 years ago (Meynard and Foch, 1992). Currently, the four-level FC-VSC is produced
by one manufacturer of industrial MV drives (Converteam, VDM6000). Up to now,
no reports are found yet on its application in the field of wind power. However, a
potential BTB-connected three- and four-level FC-VSC for a direct-drive PMSG wind
generation system can be shown in Figure 10(a) and (b), respectively. By comparing
FC-VSC with that shown in Figure 8, the FC-VSC is derived from the two-level
converter by adding dc capacitors to the series-connected switch devices. Thus, the
main features of the FC-VSC are:
.
modular structure for the switching devices; and
.
multi-level converter producing reduced dv/dt and LV harmonics.

Cell A1 Cell A1 Power cell


Power cell

Vdc/2 Vdc/2 Phase a


Ac grid Ac grid
PMSG Cell PMSG Cell
Cell B1 Cell B1
C1 Cell A2 C1 Cell A2
Cell Cell
Cell B2 Vdc/2 Transformer Cell B2
Vdc/2
Transformer
C2 C2 (necessary)
(necessary)
Neutral point Figure 9.
Wind-turbine driven
(a) (b) PMSG system based on
Notes: (a) Three-phase half-bridge GSC in each power cell; (b) single-phase full-bridge GSC five-level CHB as
machine-side converter
in each power cell
COMPEL 3-Level FC MSC 3-Level FC GSC
Ac grid
32,1

Vdc/2

46 PMSG

Vdc/2

(a) 3-level BTB FC-VSC

4-Level FC MSC 4-Level FC GSC

Ac grid

Vdc/2
PMSG
Vdc/2

Figure 10.
Wind-turbine driven
PMSG based on
BTB-connected
multi-level FC-VSC
(b) 4-level BTB FC-VSC

The FC-VSC also has some limitations, i.e.:


.
A large number of dc capacitors with separate pre-charge circuits.
.
Complex capacitor voltage balancing control. The capacitor voltages in the VSC
normally vary with the inverter operating conditions. To avoid the problems
caused by the dc voltage deviation, the voltages on the dc FCs should be tightly
controlled, which increases the complexity of the control scheme.
.
The high capacitance values and stored energies of the FCs limit the use of
FC-VSC to high switching frequency applications (e.g. $ 1,200 Hz).

C. Three-level BTB NPC-VSC for PMSG. Among the high-power converters, the
NPC inverter introduced 29 years ago (Nabae et al., 1981) is the most widely used in
all types of industrial applications, in the range of 2.3-4.16 kV, with some
applications up to 6 kV. This topology is used in a range of 3-27 MVA with
three-level topology (Rodriguez et al., 2010). Typical applications are roll mills, High-power
laminators and downhill conveyors. wind energy
Although, in the last few years, four- and five-level NPC inverters have been studied
to increase the power delivered to the load and to improve the quality of the voltage, generation
these topologies have not found an industrial application so far, due to the increased
number of clamping diodes and difficulties in dc capacitor voltage balance control.
With the development of very high-power wind turbines of over 5 MW, NPC 47
inverters can also be used in this application. As observed in Figure 11, a direct-drive
PMSG system is based on two three-level NPC-VSCs, respectively, for MSC and GSC,
which are connected BTB sharing a common dc-link.
The three-level NPC-VSC has several attractive features which lead to its
remarkable success on the market, namely:
. No dynamic voltage sharing problem. Each of the switches in the NPC converter
withstands only half of the total dc voltage during commutation.
.
Static voltage equalization without additional components. The static voltage
equalization can be achieved when the leakage current of the top and bottom
switches in an inverter leg is selected to be lower than that of the inner switches.
.
Low THD and dv/dt. The waveform of line-to-line voltages is composed of five
voltage levels, which leads to lower THD and dv/dt in comparison to the
two-level converter operating at the same voltage rating and device switching
frequency.
.
Modular design. Power-electronic building blocks on the basis of IGCTs and
IGBTs and low part count of power parts enable a modular design with excellent
reliability and availability.
.
Excellent dynamic behavior can be achieved through advanced control schemes
like VC or DTC/DPC.
.
The three-level NPC-VSC has experienced a substantial market penetrations in
2.3-4.16 kV applications that requires a low switching frequency (# 1,000 Hz)
and high converter efficiency at a lower cost.

However, there exist some drawbacks of the three-level NPC-VSC as follows:


.
Possible deviation of neutral-point voltage. It is required to control the
neutral-point potential, or the difference between both capacitors’ voltages, to
maintain a balanced and proper operation. This subject has been extensively
3-Level NPC MSC 3-Level NPC GSC

C1
Ac grid

PMSG
Figure 11.
A direct-drive PMSG
C2 system based on BTB
three-level NPC-VSCs
COMPEL studied and reported in several works over more than a decade and is considered
32,1 a solved problem in industry.
.
Unequal loss distribution. As a main structural drawbacks of the three-level
NPC-VSC, the unequal loss distribution results in asymmetrical
semiconductor-junction temperature distribution. This can be substantially
improved if active switches are placed in both NPC branches, as shown in
48 Figure 12 (three-level active NPC-VSC (Bruckner et al., 2005)), but additional 12
IGBT/diode modules are required.
.
With more than three levels, voltage balancing problems and a reduced silicon
utilization due to the increasing voltage-blocking stress of the clamping diodes
are severe structural problems of diode-clamped converters.

At present, at least two manufacturers provide BTB three-level NPC-VSCs for MV


high-power wind energy generation applications, i.e. PCS6000 by ABB (2010) and
MV7000 by Converteam (2010).

3.3 BTB CSCs


With dc inductors (chokes) as the dc energy storage, current-source inverter/converter
(CSI/CSC) technology is well suited for high-power drives (Wu et al., 2008).
The CSI can generally be divided into PWM CSI and load-commutated inverters.
The former uses symmetrical gate turn-off (GTO) or IGCT as switching devices,
whereas the latter inverter employs silicon-controlled rectifier devices.
Generally speaking, the CSCs feature a simple converter structure, low switch count,
low switching dv/dt, and reliable over-current/short-circuit protection, which are
summarized and compared with those of VSCs in Table II (Wu et al., 2008). The main
drawback of CSCs lies in its limited dynamic performance due to the use of large dc
choke. High-power CSI drives in the megawatt range are widely used in the industry
(manufactured by Allen Bradley, Power Flex7000).
The PWM CSI was proposed in the early 1970s (Phillips, 1972) and gained more
attention during the 1980s (Hombu et al., 1987). The inverter uses switching devices
with self-extinguishable capability. Prior to the advent of GCTs in the late 1990s, GTOs
were predominantly employed in high-power CSI drives. More recently, a
BTB-connected PWM CSC topology (Hombu et al., 1987) was utilized for high-power
PMSG wind power applications by Dai et al. (2007, 2009). As shown in Figure 13, the
proposed configuration consists of a direct-drive PMSG, a full-power BTB CSC, and a
3-Level Active NPC MSC 3-Level Active NPC GSC

C1 Transformer
Vdc/2 (optimal) Ac grid

Figure 12. PMSG


A direct-drive PMSG
system based on BTB
C2
three-level active
NPC-VSCs Vdc/2
High-power
Item VSC CSC
wind energy
Switch devices IGBT, GTO, IGCT with GTO, IGCT generation
anti-parallel diode
Dc-link energy storage Dc capacitor Dc choke
Converter structure Complex Simple
Switching dv/dt High Low 49
Over-current/short-circuit Difficult with IGCT Effective and reliable
protection Effective with IGBT
Dynamic performance High Low (due to low switching frequency)
PWM techniques SVM, carrier based, SHE, SVM, carrier based, and SHE Table II.
hysteresis Comparison of MV
high-power VSC
Source: Wu et al. (2008) and CSC topologies

Machine-side current source Grid-side current source


converter (MS-CSC) converter (GS-CSC)

Ldc GTO/IGCT
Transformer Ac grid
(necessary)

PMSG
Figure 13.
Cm Control schemes for the
Cg
MS-CSC and GS-CSC of a
direct-drive PMSG wind
generation system
Source: Dai et al. (2009)
grid-connected transformer. The BTB converter consists of a machine-side CSC
(MS-CSC), a grid-side CSC (GS-CSC), and filter capacitors at both sides. The MS-CSC
and GS-CSC are connected via a dc-link choke. Filter capacitors are parallel-connected
at both sides to assist current commutation as well as filter out switching harmonics.
A step-up transformer is employed to connect the converter to the grid, providing both
isolation and grid integration.
Moreover, for the time being, the integrated dc choke technology is a viable and practical
solution to the elimination of the isolation (step-up) transformer in a current-source-fed
drive (Wu et al., 2008). According to the method proposed by Wu et al. (2001), a PMSG
system based on a transformerless BTB CSC with an integrated dc choke can be shown in
Figure 14, where an integrated dc choke with a single magnetic core and four coils
MS-CSC GS-CSC

Ldc Lcm GTO/IGCT


Lg Ac grid

PMSG CM
Inductance
Figure 14.
Cm Cg PMSG based on a
transformerless BTB
CSC with an integrated
dc choke
Source: Based on Wu et al. (2008)
COMPEL (two differential coils and two common-mode (CM) coils) is used. The choke can provide
32,1 two inductances: a differential inductance Ldc that is inherently required by the CSI drive
and a CM inductance Lcm that can block the CM voltage. The use of the integrated dc choke
leads to the elimination of the isolation transformer, resulting in a significant reduction in
manufacturing and operating costs. To ensure that the stator winding of the generator is
not subject to any CM voltages, the neutral points of the grid- and machine-side filter
50 capacitors are directly connected together or through a small RL network.
In summary, the unique features of high-power MV BTB CSCs for wind-turbine
driven PMSGs are as follows:
.
simple converter structure with low switch count;
.
low switching dv/dt, fit for long-cable converter in wind-turbine driven PMSGs;
.
reliable over-current/short-circuit protection; and
.
possible transformerless design to eliminate the isolation/step-up transformers,
resulting in significant cost reduction.

However, the main drawback lies in its limited dynamic performance due to the use of
large dc choke and low switching frequency.

3.4 Matrix converters


Matrix converter is capable of converting energies between the variable ac from the
generator/motor and the constant ac of the grid in one stage. It features many advantages
which were well documented in Wheeler et al. (2002), i.e. bidirectional power flow,
sinusoidal input/output currents, and controllable input power factor. When compared
with the conventional BTB VSC, matrix converter has some significant merits, i.e.:
.
due to the absence of components with significant wear-out characteristics (such as
electrolytic capacitors), the matrix converter can be very reliable and robust; and
.
the amount of space saved by a matrix converter (no bulky energy storage
dc-link) has been estimated as a factor of three, compared with a BTB VSC.

Due to the above-mentioned features, matrix converter may become a good candidate
for wind power applications. The utilizations of a matrix converter have been explored,
respectively, with gear-drive DFIGs (Zhang et al., 1997; Cardenas et al., 2009b;
Reyes et al., 2008; Pena et al., 2009) and SCIGs (Barakati et al., 2009; Cardenas et al.,
2009b, c), and direct-drive PMSGs (Miliani et al., 2007).
For gear-drive SCIG wind generation systems with matrix converters, as shown
in Figure 15 (Barakati et al., 2009), presented a MPPT control method, named
mechanical-speed-sensorless power signal feedback, which neither requires a wind
velocity sensor nor a shaft speed sensor. While in Cardenas et al. (2009b, c), a reactive
power control scheme by estimating the input current was proposed for the
matrix-converter-supplied SCIG, with its reactive power capability investigated in detail.
On the other hand, for the direct-drive wind turbines, a system composed of a
six-phase multi-pole PMSG and a six-to-three-phase matrix converter was analyzed by
Miliani et al. (2007), which is shown in Figure 16. The presented matrix converter
operates with natural communication that makes it possible to minimize the
communication losses.
3×3 Bidirectional switches
High-power
Matrix Converter wind energy
or generation
Gear SCIG
box
51
Filter
Transformer
Ac grid
(necessary)
Figure 15.
A gear-drive SCIG
system with an indirect
matrix converter
Source: Cardenas et al. (2009a)

6×3
Matrix Converter Bidirectional switches

or

PMSG

Filter
Transformer Ac grid
(necessary) Figure 16.
A six-phase PMSG
wind turbine system
with a0 six-to-three
matrix converter
Source: Miliani et al. (2007)

Although tremendous interest has been drawn into the field of matrix converter, the use
of matrix converters in real applications is limited and the challenges (Wheeler et al.,
2002) that these applications present are very topical and important, namely:
.
The commutation problem between two controlled bidirectional devices exits, but
has been solved with the help of intelligent multi-step commutation strategies.
.
The lack of a suitably packaged bidirectional switch and a large number of
discrete power semiconductors used, has recently overcome with the
introduction of power module that includes the complete power circuit of the
matrix converter.
.
Matrix converter is not a “pure silicon converter” as expected, with passive
elements in the form of input filters required.
COMPEL .
The real challenge of the matrix converter is to be accepted in the market.
32,1 Despite that there is no intrinsic limitation to the power of a matrix converter and
megawatt range can be built, only 150 kVA matrix converter was reported to be
constructed and tested for military applications at present (Podlesak et al., 2005).

4. Control strategies for direct-drive PMSG with full-power converters


52 In spite of different topologies of VSCs for MSC and GSC, as identified in Section 3,
basic control strategies should be identical for PMSG-based wind power generation
systems. Thus, this section is to review several control schemes for a wind-turbine
driven PMSG system regardless of the power-electronic topologies of its MSC and GSC.

4.1 Classic control schemes of GSC and MSC for PMSG


The aims of control system for a PMSG wind turbine can be outlined as follows:
(1) to maximize the power extracted from the wind for a wide range of wind speeds
(known as MPPT or power optimization);
(2) to limit the power output to the rated power at high winds (power limitation);
(3) to adjust both active and reactive powers to a desirable set point according to
the operators of wind farms (power regulation); and
(4) to adjust the voltage at the grid-connected point to a desirable set point
according to modern grid codes for wind turbines (voltage regulation).
These targets can be achieved by combining power converter control for PMSGs
and blade pitch angle control for wind turbines. Generally, pitch angle control is
used during high wind speeds or under faulty grid conditions to limit active power
or current within the generator and converter ratings, which is out of scope in
this paper.
Figures 17 and 18 show a classic control scheme with alternative control modes for
the GSC and MSC of a direct-driven PMSG wind generation system according to
Chinchilla et al. (2006a, b), Ramtharan et al. (2007), Belhadj and Roboam (2007),
Michalke et al. (2007), Kim et al. (2010), Li et al. (2010) and Fernandez et al. (2010). The
VC is employed for both GSC and MSC with GVO and rotor-flux oriented, respectively.

MSC GSC
Ac grid
Isabc Lg
Igabc
PMSG C Vdc

Encoder uga ugb ugc iga igb igc


ws wr qs
SVM SVM 3s/2s 3s/2s
isa isb isc usa usb usc
Usαβ* Vgαβ Ugαβ PLL Igαβ
3s/2s 3s/2s ejqs e jq1 w q1
Ugd e–jq1 1 e–jq1
Isαβ Usαβ
Figure 17. – + + ws y f + – Ugdq Igdq
Current control schemes e–jqs e–jqs + + – – active and reactive
for PMSG’s GSC and MSC Isdq Usdq power calculations
PI PI PI w L w1 L g
PI
w s Ls w s Ls Pg , Q g
with GVO and rotor-flux active and reactive
power calculations isq∗ + – – + isq∗ igd∗+ –
1 g

– + igq∗
oriented Ps , Q s isd isq igd igq
Control mode High-power
Control mode Ugdq
Qs 0 –
wind energy

0 + 2 isd* 2 + Ug* generation
PI igq* PI
Us* 3
PI 1 + Qg*
+ 4 PI
Usdq


53
Qs
wr * – Qg
Qs +
f(ωr) PI Figure 18.
Classic control scheme
isd* igd* + Vdc* with alternative control
wr Ps* +
PI PI targets for the MSC and
– Vdc GSC of a direct-drive
– Ps
MPPT PMSG wind power
generation system
(a) Alternative control aims of MSC (b) Alternative control targets of GSC

As shown in Figure 17, the GSC and MSC can be regulated independently owing to the
decoupled capability of dc-link capacitor.
A. GSC control. As shown in Figure 18(b), in this control scheme with GVO, the
instantaneous active and reactive power outputs from GSC can be controlled by
regulating d- and q-axis currents, respectively. As a result, the GSC is capable of
maintaining the dc-link voltage constant and meanwhile accomplishing one of
following control targets, as shown in Figures 17 and 18(b):
.
Mode 1. Instantaneous reactive power control (Chinchilla et al., 2006a, b; Belhadj
and Roboam, 2007) (unity power factor favored at normal grid conditions).
.
Mode 2. Grid-voltage control (Ramtharan et al., 2007; Kim et al., 2010;
Fernandez et al., 2010), which is required according to modern grid codes for
fault-ride-through operations when the grid faults occur.

Due to the decoupling function of dc-link capacitor, the reactive power output to the
grid depends on the GSC’s rating and the generated active power, as a result, a reactive
power limiter has to be considered in practical control systems.
B. MSC control. The MSC can accomplish MPPT operation, power limitation or
power regulation if required for the PMSG-based wind power generation system.
Besides, similar to permanent magnet synchronous motor drives (Morimoto et al.,
1994), different control modes can also be applied to the MSC, namely:
. Mode 1. Minimum stator current control (Belhadj and Roboam, 2007;
Michalke et al., 2007), where d-axis stator current is controlled to be null.
However, this control mode requires increasing the MSC power rating since the
generator absorbs the reactive power from the converter.
. Mode 2. Unity power factor control (Michalke et al., 2007), in which some d-axis
stator current should be injected so as to compensate the reactive power
absorbed by the PMSG. This control mode minimizes the MSC power rating.
Whereas, as the stator voltage is not directly regulated and thus varies with the
generator speed, this control mode could cause over-voltages for the generator
and the converter in the case of over speeds. Moreover, since the reactive power
COMPEL output to the grid is independent from the generator reactive power, it is less
32,1 necessary to control the reactive power at the generator side.
.
Mode 3. Constant stator-voltage control (Ramtharan et al., 2007; Belhadj and
Roboam, 2007; Michalke et al., 2007), where the stator voltage is controlled
instead of the reactive power to avoid the risk of over-voltage due to over speeds.
In this case, the generator and converter operate at the rated voltage for which
54 they are designed and optimized through entire speed range. However, this
control mode also requires increased MSC power rating, due to the reactive
power absorbed by the generator.
.
Mode 4. Maximum efficiency control (Chinchilla et al., 2006a), in which the
generator reactive power demand is calculated and imposed in order to minimize
power losses, both in the generator (including copper loss and iron loss
(Morimoto et al., 1994; Nakamura et al., 2002) and power converter, along the
whole operating range. Apparently, the MSC power rating needs to be increased
as well.

As long as the MSC’s power rating permits, modes 4 and 3 are suggested to be used
below and above the rated generator speed, respectively, via seamless switches.

4.2 Active damping control


As the drive train of a wind turbine behaves as a torsional spring characteristic
(Akhmatov, 2003), the generator speed is usually prone to oscillations once the
system gets excited by electrical or mechanical load changes (Hansen and
Michalke, 2008, 2009; Michalke and Hansen, 2010). The oscillation leads to the
fluctuation in the output power and the increase of mechanical fatigue of the drive
train. Moreover, typical frequency of the oscillation is as low as 0.1-10 Hz, which tends
to coincide with the frequencies associated with power system inter-area oscillations
(0.1-2.5 Hz) (Geng et al., 2010). As a consequence, instability issues are expected for the
power system with high wind power penetrations if there is no damping technique
adopted for wind turbines. In the traditional power plant, the wound-field synchronous
generators are directly connected to the grid. The damper winding in the rotor
can provide additional damping torque to alleviate speed oscillations. However, this
benefit does not hold true for a PMSG-based wind turbine with full-power converter
connected to the grid, as this configuration does not even have a damper winding.
Hence, it is necessary to use external damping system to suppress the oscillations
and thus avoid instability. A popular way is to mount the damping rubber on the
drive train shaft. However, it is costly and requires additional installation space on
the drive train shaft.
The other way is to apply active damping techniques either on the generator side
or on the grid side. Since the output power is expected to oscillate if the generator
mechanical resonance exists, positive damping is achieved if the GSC performs as a
damper for the fluctuating power. This strategy seems effective as the grid has
almost infinite inertia. However, the fluctuating power transferred to the grid is
adverse to the power system stability. Thus, this strategy is usually used in the
situation with low wind power penetrations. Otherwise, additional circuits such as
the STATCOM are necessary to consume the oscillated power (El-Moursi et al., 2010).
Alternatively, active damping can be applied from the generator side (Hansen and
Michalke, 2008, 2009) or with dc-link current estimation between MSC and GSC High-power
(Geng et al., 2010). wind energy
According to Hansen and Michalke (2008), Figure 19 shows the control scheme with
active damping controller for the MSC and GSC of a wind-turbine driven PMSG. As generation
shown in Figure 19, this control scheme employs SVO and GVO VCs, respectively, for
the MSC and GSC. As shown, it consists of the following controllers:
.
The active damping controller ensures a stable operation of the wind turbine 55
system, by damping the torsional oscillations excited by the drive train and
reflected in the generator speed vr.
.
The MSC controller maintains the dc-link voltage constant (which, however, is
kept constant by GSC in Figure 18) and controls the stator voltage of generator to
its rated value in the SVO reference frame throughout the entire operations.
.
The GSC controller controls the instantaneous active power (MPPT operation,
which is accomplished by MSC in Figure 18) and reactive power output to the
grid in the GVO reference frame.

The control of the active power and the dc-link voltage is strongly related to each other,
as the active power can be fed into the grid as long as the dc-link voltage is maintained
constant. Thus, in this control scheme, the MPPT is implemented by GSC while the
dc-link voltage is kept constant by MSC, which is opposite to the classic control scheme
shown in Figure 18.
In terms of active damping controller, it consists of a band-pass filter and a phase
compensator. The input of the damping controller is the generator speed, while the
output is an additional term added to the desired dc-link voltage reference. As a
result, the dc-link voltage is controlled to its set value Vdc * þ DVdc *, which is provided
by the damping controller and oscillates with the right frequency and phase
angle, generating a torque that mitigates the speed oscillations. It is worth pointing out
that even though the aim is to maintain the dc-link voltage constant and ensure the
power to be transferred from the PMSG terminals to the grid, small voltage variations
of the dc-link capacitor are allowed, as the electrical damping of the system is
necessary.

Usdq
Qg –
– isq* igq* + Qg *
Us*
PI PI

isd* MPPT
Phase ∆Vdc + + igd*
+ Pg* Figure 19.
compensator PI
Vdc* + – A modified control scheme
wr Vdc Pg – with active damping
Bandpas
wr controller for MSC and
s filter Damping controller
GSC of a PMSG-based
(a) Control aim of MSC (b) Control target of GSC wind power
generation system
Source: Hansen and Michalke (2008)
COMPEL More recently, an active damping control strategy with dc-link current estimation
32,1 was presented for a PMSG-based wind power generation system in Geng et al. (2010)).
According to the description in Geng et al. (2010), the control scheme can be shown in
Figure 20. As shown, GSC is responsible for keeping dc-link voltage constant and
realizing unity power factor, while MSC accomplishes the MPPT operation and active
damping control based on dc-link current estimations.
56
4.3 Sensorless control
The absence of gearbox in the direct-drive PMSG-based wind turbines increases the
reliability and decreases the maintenance requirements. It is well known that shaft
encoders are fragile devices and its use implies extra wiring and breakable mechanical
mounting which may deteriorate the inherent robustness of the PMSG system. To
further enhance the robustness in the entire wind power generation system a sensorless
control scheme can be used, so that the problems related with electromagnetic
interferences in the position signals and failures in the position sensor can be avoided.
As VSCs are employed as machine-side converters of PMSG wind generation
systems, as shown in Figure 17, all the sensorless control strategies in PMSM drives
can be transplanted and applied in the wind-turbine PMSGs if necessary. Generally,
the sensorless control strategies in PMSM drives can be classified into two big families
( Johnson et al., 1999, p. 1) fundamental excitation method and saliency and signal
injection method. In the saliency and signal injection method, the difference of saliency
between direct and quadrature axes is utilized to estimate the rotor position. A suitable
signal such as high-frequency voltage or current component is injected from the
inverter in order to detect inductance variation with the rotor position (Ribeiro et al.,
1998; Hyunbae and Lorenz, 2004; Jang et al., 2003; Holtz, 2008; Li et al., 2009). The rotor
position can be precisely estimated at very low speeds and even standstill by this
method. However, it is quite complicated and requires great attention in the design of
the control system and of the signal-conditioning system to extract the rotor position
and speed. Further, because unwanted torque and additional losses may be generated
by the injected signal, the methods based on saliency and signal injection are mostly
used at standstill and low speeds. On the other hands, the fundamental excitation

Qs
Qs* + – isd* Vdc –
PI igd* +Vdc*
Ssabc Sgabc
PI
lsabc lgabc

DC-link – Ps Te – isq*
current Ps* +
PI Qg
Figure 20. –
estimator + + igq* +Qg*
Control scheme for the + + PI
MSC and GSC of a idc Vdc kp
PI Te_com
wind-turbine driven Tewr – Pg +
PMSG with active Damping controller
damping controller based
on dc-link current (a) Control aim of MSC (b) Control target of GSC
estimation
Source: According to Geng et al. (2010)
method does not need any additional test signals and estimates the rotor position and High-power
speed from the fundamental components of stator voltages and currents. It is hard to wind energy
estimate position at the low-speed region, which, however, is not a severe issue in the
application of wind-turbine driven PMSGs, as wind power generation systems seldom generation
operate at very low speeds. Thus, it can be concluded that sensorless schemes based on
the fundamental excitation method are superior to those based on the saliency and signal
injection method in the application of PMSG-based wind power generation systems. 57
In the following section, two popular sensorless schemes, respectively, based on
phase-locked loop (PLL) technique and sliding mode observer (SMO) are reviewed.
A. PLL technique. Among various sensorless schemes based on the fundamental
excitation method, a PLL technique, which is widely used in grid-connected
converters, is favored to capture the rotor speed and to synchronize the estimated d-q
frame to the actual rotor-flux frame in the PMSM drives (Harnefors and Nee, 2000;
Harnefors et al., 2003; Morimoto et al., 2002; Burgos et al., 2006; Li et al., 2007), and
PMSG wind generation systems (Kawabe et al., 2007; Inoue et al., 2008; Diaz et al., 2009;
Hu and Xu, 2009). Although this concept was presented in Harnefors and Nee (2000),
Harnefors et al. (2003) and Morimoto et al. (2002)), applied in the PMSM drives, it has
not been exploited recently (Comanescu and Xu, 2006), where a simple d-q frame PLL
presented in Kawabe et al. (2007) was used to estimate the stator frequency of an
induction machine by aligning the PLL to its terminal voltages.
Most of these PLL-based sensorless schemes above-mentioned for both PMSMs and
PMSGs are implemented by the estimation of an extended electromotive force (EMF)
(Morimoto et al., 2002; Chen et al., 2003), as:

2 3 2 3 2 3
e^ sd ~ ^isq
  2sin us
4 5 ¼ vs ½ðLsd 2 Lsq Þisd þ cf  2 ðLsd 2 Lsq Þð pisq Þ 4 5 þ ðv^s 2 vs ÞLsd 4 5
e^ sq cos u~s ^isd

ð1Þ

where e^ sd and e^ sq are the estimated d- and q-axis extended EMFs, Lsd and Lsq are the
respective d- and q-axis inductances, isd and isq are the respective d- and q-axis stator
currents in the actual synchronous dq reference frame, ^isd and ^isq are the respective d-
and q-axis stator currents in the estimated synchronous d^ q^ reference frame, vs and v^s
are the rotating angular speeds of the actual and estimated synchronous frames,
respectively, and u^s is the spatial angle by which the estimated synchronous frame lags
the actual one.
As indicated in Morimoto et al. (2002), the model given in equation (1) can be utilized
by any type of synchronous machines such as surface PMSMs (Lsd ¼ Lsq), interior
PMSMs (Lsd , Lsq) and synchronous reluctance machines (cf ¼ 0). Further, as shown
in equation (1), the extended EMF in the estimated d^ q^ synchronous reference frame
includes the information of estimation position error u~s , which can be utilized for the
estimation of rotor position and speed. The extended EMF e^ sd and e^ sq in the d^ q^ frame
can be estimated by different methods, such as, least-order observer in Morimoto et al.
(2002) and Kawabe et al. (2007) disturbance observer in Chen et al. (2003) and directly
calculated based on voltage model with current variations neglected in Li et al. (2007)
and Hu and Xu (2009)).
COMPEL Assuming that the error between the estimated speed and the actual speed is
32,1 sufficiently small, equation (1) can be simplified as:
" # " #
e^ sd   2sin u~s
¼ vs ½ðLsd 2 Lsq Þisd þ cf  2 ðLsd 2 Lsq Þð pisq Þ ð2Þ
e^ sq cos u~s
58
Once the extended EMF are obtained, the estimated position error is derived by:
 
~ e^ sd
us ¼ 2arctan ð3Þ
e^ sq
Consequently, the obtained estimated position error u~s is regulated to be zero by PLL
scheme so that the estimated d^ q^ frame is synchronized to the actual dq rotor frame.
Based on the analysis aforementioned, the structure of PLL scheme based on extended
EMF for the estimation of rotor position and speed is shown in Figure 21.
It can be seen from the voltage model that the estimation of extended EMF usually
employs machine’s parameters, namely, Rs, Lsd and Lsq, thus, the variations of
these parameters, especially Lsq in IPMSM due to magnetic saturation, will deteriorate
the estimated accuracy. Online parameter identification method can be used to
identify these machine’s parameters (Morimoto et al., 2006). Moreover, the dq-axis
cross-coupling magnetic saturation may also affect the accuracy of the rotor position
and speed estimation in the sensorless schemes based on the extended EMF,
which should be considered to improve the precision in the system control design
(Li et al., 2007).
Furthermore, as an extension or generalization to extended EMF (Morimoto et al.,
2002; Burgos et al., 2006) and fictitious PM flux (Koonlaboon and Sangwongwanich,
2005) concepts, the “active flux” concept was proposed by Boldea et al. (2008), which, as
a result turns all rotor salient-pole ac machines into fictitious non-salient-rotor
machines such that the rotor position and speed estimations become simpler and
applicable to all ac machines.
Without cross-coupling saturation considered, the active flux ca is defined as the
flux that multiplies the q-axis current iq in the dq-model torque expression of all ac
machines in the rotor synchronous reference frame. For all types of PM synchronous
machines:

PLL PLL ŵs


ŵs
0 + ki 1 0+ k 1
kp + qˆs kp + i s qˆs
Figure 21. s s s
– –
Schematic diagram of
rotor speed and position qs
~
êsd
estimator based on PLL êsd, êsq Extended EMF Extended EMF
and extended EMF (3) Observer
Observer
according to (a) Morimoto
Isab Usab
et al. (2002) and (b) Burgos Isab Usab
et al. (2006)
(a) (b)
3 High-power
Te ¼ pjc a jisq ð4Þ
2 wind energy
with:
generation
jc a j ¼ cf þ ðLsd 2 Lsq Þisd ð5Þ
where the d-axis is aligned with the active flux (active flux orientation), namely,
ca ¼ cad þ j 0, and still corresponds to the rotor pole (PM) axis, namely, ua ¼ us, 59
as shown in Figure 22. As shown, the active flux position is identical to the rotor position
in any operation mode, which leads to a significant simplification in rather accurate
estimation of rotor position and speed whenever the active flux estimation is feasible.
Besides, the stator-voltage model can be given in the stator stationary reference
frame as:
U ss ¼ Rs I ss þ d css =dt ð6Þ
where superscript s refers to stator stationary reference frame.
Thus, the active flux observer csa in the stator stationary reference frame can be
derived as:
Z
 s 
cas ¼ U s 2 Rs I ss þ DU ss dt 2 Lsq I ss ð7Þ

where DU ss compensates various errors in the cs estimation including inverter


nonlinearity, dead time, integration dc-offset, and stator resistance variation, and:
cas ¼ ca cos ua þ jca sin ua ð8Þ
Once the active flux is estimated based on equation (7), a PLL technique can be used to
track the rotor position and speed precisely by regulating caq to be null, as shown in
Figure 23. As shown, a compensation block is utilized and a measured q-axis
inductance as a function of q-axis current or torque is inserted for improving the
accuracy of the active flux estimation. It is reported from Boldea et al. (2009) and
Paicu et al. (2009)) that by using the sensorless scheme shown in Figure 23, the vector
controlled and direct torque- and flux-controlled IPMSMs can both operate at the
lowest speed limit of 2 rpm with half-rated load torque.
B. Sliding mode observer. An adaptive SMO was initially presented to obtain a
position-and-velocity sensorless control for brushless dc motors by Furuhashi et al.
(1992) and Chen et al. (2000). Detailed design process of the adaptive observer can be
found in Chen et al. (2000) and the scheme of the adaptive SMO is shown in Figure 24.

q F
ws d
Figure 22.
Spatial relationship
between the stationary
yf + (Ld – Lq)id frame and synchronous
dq frame aligned with
qa
the “active flux”
a
COMPEL Stator flux observer
Isab "Active flux"
32,1 Lsq (isq)

Rs PLL ŵs
– – ŷaq
Usab + yˆsab + yˆsa ki 1
ˆ kp +
1/s e–jθs s s
60 + – qˆs
∆Uss kic
kpc +
s
+
Î sdq yˆsdq* yˆsab*
Figure 23. ˆ Current ˆ

Active flux observer and e–jqs Model e jqs


rotor position/speed
estimator based on Compensation block
PLL scheme
Source: Boldea et al. (2008)

Isab

Isab
B1 ˆ
I sab
0 + Iˆsab +
Eˆsab
1 –
[I 0]
s

[0 I]
A11 A12
Eˆsab
0 Aˆ22 qˆs
arctan ( )
Adaptive wˆ s
LPF
scheme
Figure 24. Velocity estimation
Structure of the
K sgn (Isab - Î sab)
adaptive sliding
observer for SPMSM
Source: Chen et al. (2000)

As shown, since the position and velocity estimations are integrated into one observer,
rotor position estimation error may occur when velocity estimation error exits. Thus,
robustness of the position estimation to the velocity estimation error cannot be
neglected and an additional adaptive scheme was designed for velocity estimation
precisely under both steady-state and transient conditions.
Further, the rotor position and velocity observers shown in Figure 24 are
only feasible for surface-mounted PMSM, while for IPMSM system, extended EMF
defined in equation (2) can be adopted (Xu and Rahman, 2007a). The scheme of
the adaptive SMO for IPMSM is shown in Figure 25. As shown, a simplified
Kalman filter (Xu and Rahman, 2007b) was used to estimate the rotor speed, which is
different from that shown in Figure 24. Although, this adaptive SMO was designed
for IPMSM drives, it can also be applied in other types of synchronous machines,
Isab High-power
Usab B1
wind energy
Iˆsab generation
0 Iˆsab
+ Eˆsab +
1 –
s [I 0]
61
[0 I ]
A11 A12
Eˆsab
0 Aˆ22 qs
ˆ

arctan ( )
ω̂ s Simplified Kalman
Filter
Velocity estimation Figure 25.
Structure of the adaptive
K sgn (Isab - Î sab) sliding observer
for IPMSM
Source: Xu and Rahman (2007a)

such as surface PMSMs (Lsd ¼ Lsq) (Xu et al., 2009) and synchronous reluctance
machines (cf ¼ 0).
For the purpose of improving the accuracy of the rotor position and velocity, a
measured q-axis inductance as a function of q-axis current should be employed in the
observer. It is reported from Xu and Rahman (2007b) that by using the sensorless
scheme shown in Figure 25, the direct torque controlled IPMSM can operate at low
speed down to 10 rpm with half-rated load torque.
However, both observers shown in Figures 24 and 25 are closed-loop speed
adaptive. The speed adaption is usually performed at the last step of the estimation
process. Hence, the speed estimate is affected by cumulative errors, noise and delays.
When the inaccurately estimated speed value is fed back to the observer, the stator
flux, rotor position and speed estimations gradually worsen. This may easily lead the
drive to instability, particularly, at low speeds.
With active flux given in equation (5) introduced, a stator-flux observer was
proposed in Foo and Rahman (2010a, b). The observer is inherently sensorless and does
not require any speed adaption mechanism. According to Foo and Rahman (2010a), the
observer for estimating stator flux, rotor position and speed can be shown in Figure 26.

Isab – ŷaab
Lsq (isq) arc tan ()
+ Lsq (isq)
Rs q̂s
Usab – ŷsab ŷsdq
+
1 s ˆ Current
e–jqs Figure 26.
+ model
Structure of the stator flux
K Iˆsdq and rotor position
Iˆsab observer presented in Foo
ˆ and Rahman (2010a, b)
ejqs
COMPEL As shown, the observer utilizes two reference frames, namely, stator stationary ab
32,1 frame and rotor synchronous dq frame. The stator flux is estimated in the ab frame,
and thus the active flux and rotor position can be calculated. While, the stator current
is estimated with current model in the rotor dq frame. The current estimation errors
are then used as feedback signals through a gain K. The stator-flux observer combines
the advantages of the current model at low speeds and those of the voltage model at
62 higher speeds.
It is worth noting that the observer synthesis in both the ab frame and the rotor
dq frame is only possible by introducing the active flux (Boldea et al., 2008) as it
provides the rotor position information. Moreover, the rotor speed information is
absent from the observer model owing to the use of the two reference frames. As a
result, unlike the conventional speed adaptive observers (Figures 24 and 25), any errors
associated with inaccurate speed estimation is removed, which improves the
stator-flux estimation at very low speeds and standstill. It is reported from Foo and
Rahman (2010a, b) that with VC and DTC schemes, respectively, the sensorless
controlled IPMSM by using the observer from Figure 26 can operate at very low speeds
including standstill and 5 rpm without signal injection. It is worthwhile to indicate that
the both measured q-axis inductance as a function of q-axis current and the estimated
stator resistance are required to be employed in the observer and control system to
improve the accuracy.

5. Conclusions
Power-electronic systems have been playing a significant role in the integration of
large-scale wind turbines into power systems due to the fact that during the past three
decades power-electronic technology has experienced a dramatic evolution, including
advents of high-power-handled semiconductor devices, a variety of grid-friendly
converters, and easily practical implementations of advanced control strategies as well
as PWM techniques by using extremely high-speed DSPs.
For high-power PMSG-based wind turbines ranging from 3 to 5 MW,
parallel-connected two-level LV BTB VSCs are the most cost-effective converter
topology with mature commercial products, particularly for dual stator-winding PMSG
generation systems. For higher-capacity wind-turbine driven PMSGs rated from 5 to
10 MW, MV multi-level converters, such as five-level regenerative CHB, three- and
four-level FC BTB VSC, and three-level BTB NPC-VSC, are preferred. Among them,
three-level BTB NPC topology is the favorite with well-proven technology and
industrial applications, which can also be extensively applicable with open-end
winding and dual stator-winding PMSGs so as to create even higher voltage/power
wind generation systems. As a counterpart to multi-level VSCs, GTO/IGCT-based
CSCs feature a simple converter structure, but are limited by its relatively poor
dynamic performance.
As to the control systems, first for the partial-converter-supplied gear-drive DFIGs,
advanced control strategies, including FOC and DPC, which require no synchronous
transformations, keep well steady-state and transient responses and feature strong
robustness, are preferred. Moreover, improved control schemes and protection
measures are expected to enhance the fault-ride-through capability during grid-voltage
disturbances. While for the direct-drive PMSGs, active damping control techniques
should be taken into account in the basic control schemes of machine-side converters
and GSCs, otherwise, instability issues are expected for the power system particularly High-power
with high penetrations of wind power. Further, sensorless control algorithms, wind energy
such as PLL technique and SMO both based on fundamental components, are
suggested to be employed for enhancing the robustness in the entire PMSG-based wind generation
power generation system, due to that the problems related with electromagnetic
interferences in the position signals and the failures in the mechanical encoders can
be avoided. 63

References
Abad, G., Rodriguez, M.A. and Poza, J. (2008a), “Two-level VSC-based predictive direct
power control of the doubly fed induction machine with reduced power ripple at low
constant switching frequency”, IEEE Trans. on Energy Conversion, Vol. 23 No. 2,
pp. 570-80.
Abad, G., Rodriguez, M.A. and Poza, J. (2008b), “Two-level VSC-based predictive torque
control of the doubly fed induction machine with reduced torque and flux ripples
at low constant switching frequency”, IEEE Trans. on Power Electronics, Vol. 23 No. 3,
pp. 1050-61.
ABB (2010), PCS 6000 Wind: Medium Voltage Full Power Frequency Converter for Wind
Turbines, available at: www05.abb.com/global/scot/scot232.nsf/veritydisplay/
62b06de8d8d9b942c125765f002dc31f/$File/PCS%206000%20Wind%20Brochure.pdf
(accessed 30 September).
Akhmatov, V. (2003), “Analysis of dynamic behavior of electric power systems with large
amount of wind power”, PhD dissertation, Orsted-DTU, Technical University of Denmark,
Copenhagen.
Barakati, S.M., Kazerani, M. and Aplevich, J.D. (2009), “Maximum power tracking control for a
wind turbine system including a matrix converter”, IEEE Trans. on Energy Conversion,
Vol. 24 No. 3, pp. 705-13.
Baroudi, J.A., Dinavahi, V. and Knight, A.M. (2005), “A review of power converter topologies for
wind generators”, Proc. 2005 IEEE International Conference on Electric Machines and
Drives, San Antonio, TX, pp. 458-65.
Belhadj, J. and Roboam, X. (2007), “Investigation of different methods to control a small
variable-speed wind turbine with PMSM drives”, Journal of Energy Resources Technology,
Vol. 129 No. 3, pp. 200-13.
Beltran, B., Ahmed-Ali, T. and Benbouzid, M.E.H. (2008), “Sliding mode power control of
variable-speed wind energy conversion systems”, IEEE Trans. on Energy Conversion,
Vol. 23 No. 2, pp. 551-8.
Boldea, I., Paicu, M.C. and Andreescu, G.D. (2008), “Active flux concept for motion-sensorless
unified AC drives”, IEEE Trans. on Power Electronics, Vol. 23 No. 5, pp. 2612-18.
Boldea, I., Paicu, M.C., Andreescu, G.D. and Blaabjerg, F. (2009), “‘Active flux’ DTFC-SVM
sensorless control of IPMSM”, IEEE Trans. on Energy Conversion, Vol. 24 No. 2,
pp. 314-22.
Bruckner, T., Bernet, S. and Guldner, H. (2005), “The active NPC converter and its loss-balancing
control”, IEEE Trans. on Industrial Electronics, Vol. 52 No. 3, pp. 855-68.
Bueno, E.J., Cobreces, S., Rodriguez, F.J., Hernandez, A. and Espinosa, F. (2008), “Design of a
back-to-back NPC converter interface for wind turbines with squirrel-cage induction
generator”, IEEE Trans. on Energy Conversion, Vol. 23 No. 3, pp. 932-45.
COMPEL Burgos, R.P., Kshirsagar, P., Lidozzi, A., Jang, J., Wang, F., Boroyevich, D., Rodriguez, P. and
Sul, S.K. (2006), “Design and evaluation of a PLL-based position controller for sensorless
32,1 vector control of permanent-magnet synchronous machines”, Proc. 32nd Annual
Conference on IEEE Industrial Electronics (IECON 2006), pp. 5081-6.
Cardenas, R., Pena, R., Wheeler, P. and Clare, J. (2009a), “Control of the reactive power supplied
by a matrix converter”, IEEE Trans. on Energy Conversion, Vol. 24 No. 1, pp. 301-3.
64 Cardenas, R., Pena, R., Wheeler, P., Clare, J. and Asher, G. (2009b), “Control of the reactive power
supplied by a WECS based on an induction generator fed by a matrix converter”, IEEE
Trans. on Industrial Electronics, Vol. 56 No. 2, pp. 429-38.
Cardenas, R., Pena, R., Tobar, G., Clare, J., Wheeler, P. and Asher, G. (2009c), “Stability
analysis of a wind energy conversion system based on a doubly fed induction generator
fed by a matrix converter”, IEEE Trans. on Industrial Electronics, Vol. 56 No. 10,
pp. 4194-206.
Carrasco, J.M., Franquelo, L.G., Bialasiewicz, J.T., Galvan, E., Guisado, R.C.P., Prats, M.A.M.,
Leon, J.I. and Moreno-Alfonso, N. (2006), “Power-electronic systems for the grid integration
of renewable energy sources: a survey”, IEEE Trans. on Industrial Electronics, Vol. 53
No. 4, pp. 1002-16.
Chen, Z. and Spooner, E. (1998), “Grid interface options for variable-speed permanent-magnet
generators”, Proc. IEE EPA, Vol. 145 No. 4, pp. 273-83.
Chen, Z. and Spooner, E. (2003), “Current source thyristor inverter and its active compensation
system”, Proc. IEE GTD, Vol. 150 No. 4, pp. 447-54.
Chen, Z., Guerrero, J.M. and Blaabjerg, F. (2009), “A review of the state of the art of power
electronics for wind turbines”, IEEE Trans. on Power Electronics, Vol. 24 No. 8, pp. 1859-75.
Chen, Z., Tomita, M., Doki, S. and Okuma, S. (2000), “New adaptive sliding observers for position-
and velocity-sensorless controls of brushless DC motors”, IEEE Trans. on Industrial
Electronics, Vol. 47 No. 3, pp. 582-91.
Chen, Z., Tomita, M., Doki, S. and Okuma, S. (2003), “An extended electromotive force model for
sensorless control of interior permanent-magnet synchronous motors”, IEEE Trans. on
Industrial Electronics, Vol. 50 No. 2, pp. 288-95.
Chinchilla, M., Arnaltes, S. and Burgos, J.C. (2006a), “Control of permanent-magnet generators
applied to variable-speed wind-energy systems connected to the grid”, IEEE Trans. on
Energy Conversion, Vol. 21 No. 1, pp. 130-5.
Chinchilla, M., Arnaltes, S., Burgos, J.C. and Rodrı́guez, J.L. (2006b), “Power limits of
grid-connected modern wind energy systems”, Renewable Energy, Vol. 31 No. 9, pp. 1455-70.
Choi, J.W. and Sul, S.K. (1998), “Fast current controller in three-phase AC/DC boost
converter using d-q axis cross coupling”, IEEE Trans. on Power Electronics, Vol. 19 No. 1,
pp. 179-85.
Codd, I. (2003), “Windfarm power quality monitoring and output comparison with EN50160”,
Proc. of the 4th Intern. Workshop on Large-scale Integration of Wind Power and
Transmission Networks for Offshore Wind Farm, Sweden.
Comanescu, M. and Xu, L. (2006), “An improved flux observer based on PLL frequency estimator
for sensorless vector control of induction motors”, IEEE Trans. on Industrial Electronics,
Vol. 53 No. 1, pp. 50-6.
Converteam (2010), “MV7000: entering a new dimension for reliability and performance in
medium-voltage AC drives”, available at: www.converteam.com/converteam/1/doc/News/
070712_Converteam_MV7000.pdf (accessed 30 September).
Dai, J., Xu, D. and Wu, B. (2007), “A novel control system for current source converter based High-power
variable speed PM wind power generators”, IEEE Power Electronics Specialists Conference
(PESC 2007), Orlando, FL, pp. 1852-7.
wind energy
Dai, J., Xu, D. and Wu, B. (2009), “A novel control scheme for current-source-converter-based generation
PMSG wind energy conversion systems”, IEEE Trans. on Power Electronics, Vol. 24 No. 4,
pp. 963-72.
Datta, R. and Ranganathan, V.T. (2001), “Direct power control of grid-connected wound rotor 65
induction machine without rotor position sensors”, IEEE Trans. on Power Electronics,
Vol. 16 No. 3, pp. 390-9.
Diaz, S.A., Silva, C., Juliet, J. and Miranda, H.A. (2009), “Indirect sensorless speed control of a
PMSG for wind application”, Proc. IEEE International Electric Machines and Drives
Conference (IEMDC 2009), pp. 1844-50.
El-Moursi, M.S., Bak-Jensen, B. and Abdel-Rahman, M.H. (2010), “Novel STATCOM controller
for mitigating SSR and damping power system oscillations in a series compensated wind
park”, IEEE Trans. on Power Electronics, Vol. 25 No. 2, pp. 429-41.
Enercon GmbH (2009), Enercon Wind Turbines: Technology and Service, Aurich, available at:
www.enercon.de/en/_home.htm (accessed 31 March 2010).
Fernandez, L.M., Garcia, C.A. and Jurado, F. (2010), “Operating capability as a PQ/PV node of a
direct-drive wind turbine based on a permanent magnet synchronous generator”,
Renewable Energy, Vol. 35 No. 6, pp. 1308-18.
Foo, G. and Rahman, M.F. (2010a), “Sensorless direct torque and flux-controlled IPM
synchronous motor drive at very low speed without signal injection”, IEEE Trans. on
Industrial Electronics, Vol. 57 No. 1, pp. 395-403.
Foo, G. and Rahman, M.F. (2010b), “Sensorless vector control of interior permanent magnet
synchronous motor drives at very low speed without signal injection”, IET Electr. Power
Appl., Vol. 4 No. 3, pp. 131-9.
Furuhashi, T., Sangwongwanich, S. and Okuma, S. (1992), “A position-and-velocity sensorless
control for brushless DC motors using an adaptive sliding mode observer”, IEEE Trans.
on Industrial Electronoics, Vol. 39 No. 2, pp. 89-95.
GE (2010), Low Voltage Ride Through Technology, available at: www.gepower.com/businesses/
ge_wind_energy/en/downloads/ge_lvrt_brochure.pdf (accessed 30 September).
Geng, H., Xu, D., Wu, B. and Yang, G. (2010), “Active damping for PMSG based WECS with DC
link current estimation”, IEEE Trans. on Industrial Electronics, Vol. 58 No. 4, pp. 1110-19.
Hansen, A.D. and Michalke, G. (2008), “Modelling and control of variable-speed multi-pole permanent
magnet synchronous generator wind turbine”, Wind Energy, Vol. 11 No. 5, pp. 537-54.
Hansen, A.D. and Michalke, G. (2009), “Multi-pole permanent magnet synchronous generator
wind turbines’ grid support capability in uninterrupted operation during grid faults”, IET
Renewable Power Generation, Vol. 3 No. 3, pp. 333-48.
Hansen, A.D., Michalke, G., Sorensen, P. and Lund, T. (2006), “Coordinated voltage control of
DFIG wind turbines in uninterrupted operation during grid faults”, Wind Energy, Vol. 10
No. 10, pp. 51-68.
Haque, M.E., Muttaqi, K.M. and Negnevitsky, M. (2008a), “Control of a standalone variable speed
wind turbine with a permanent magnet synchronous generator”, Proc. IEEE Power and
Energy Society General Meeting – Conversion and Delivery of Electrical Energy in the 21st
Century, Pittsburgh, PA, pp. 1-9.
COMPEL Haque, M.E., Negnevitsky, M. and Muttaqi, K.M. (2008b), “A novel control strategy for a variable
speed wind turbine with a permanent magnet synchronous generator”, Proc. IEEE
32,1 Industry Applications Society Annual Meeting (IAS 2008), Edmonton, Canada, pp. 1-8.
Haque, M.E., Negnevitsky, M. and Muttaqi, K.M. (2010), “A novel control strategy for a
variable-speed wind turbine with a permanent-magnet synchronous generator”, IEEE
Trans. on Industry Applications, Vol. 46 No. 1, pp. 331-9.
66 Harnefors, L. and Nee, H.P. (2000), “A general algorithm for speed and position estimation of AC
motors”, IEEE Trans. on Industrial Electronics, Vol. 47 No. 1, pp. 77-83.
Harnefors, L., Jansson, M., Ottersten, R. and Pietilainen, K. (2003), “Unified sensorless vector
control of synchronous and induction motors”, IEEE Trans. on Industrial Electronics,
Vol. 50 No. 1, pp. 153-60.
He, Y., Hu, J. and Zhao, R. (2005), “Modeling and control of wind-turbine used DFIG under
network fault conditions”, Proc. of the 8th International Conference on Electrical Machines
and Systems, ICEMS2008, Nanjing.
Higuchi, Y., Yamamura, N., Ishida, M. and Hori, T. (2000), “An improvement of performance for
small-scaled wind power generating system with permanent magnet type synchronous
generator”, Proc. 26th Annual Conference of the IEEE Industrial Electronics Society
(IECON 2000), Nagoya, Japan, Vol. 2, pp. 1037-43.
Holtz, J. (2008), “Acquisition of position error and magnet polarity for sensorless control of
PM synchronous machines”, IEEE Trans. on Industry Applications, Vol. 44 No. 4,
pp. 1172-80.
Hombu, M., Ueda, S. and Ueda, A. (1987), “A current source GTO inverter with sinusoidal inputs
and outputs”, IEEE Trans. on Industry Applications, Vol. IA-23 No. 2, pp. 247-55.
Hu, J. and He, Y. (2009), “Reinforced control and operation of DFIG-based wind-power-generation
system under unbalanced grid voltage conditions”, IEEE Trans. on Energy Conversion,
Vol. 24 No. 4, pp. 905-15.
Hu, J., He, Y. and Xu, L. (2008a), “Dynamic modeling and direct power control of wind turbine
driven DFIG under unbalanced network voltage conditions”, J. Zhejiang Univ. Sci. A, Vol. 9
No. 12, pp. 1731-40.
Hu, J., He, Y., Xu, L. and Williams, B.W. (2009), “Improved control of DFIG systems during
network unbalance using PI-R current regulators”, IEEE Trans. on Industrial Electronics,
Vol. 56 No. 2, pp. 439-59.
Hu, J., Shang, L., He, Y. and Zhu, Z.Q. (2011), “Direct active and reactive power regulation of
grid-connected DC/AC converters using sliding mode control approach”, IEEE Trans. on
Power Electronics, Vol. 26 No. 1, pp. 210-22.
Hu, J., Nian, H., Hu, B., He, Y. and Zhu, Z.Q. (2010), “Direct active and reactive power regulation of
DFIG using sliding mode control approach”, IEEE Trans. on Energy Conversion, Vol. 25
No. 4, pp. 1028-39.
Hu, S. and Xu, H. (2009), “Research on sensorless control based back-to-back converter for
direct-driven WECS”, 2009 Asia-Pacific Power and Energy Engineering Conference
(APPEEC 2009), pp. 1-4.
Hu, W., Wang, Y., Yao, W., Zhang, H., Wu, J. and Wang, Z. (2008b), “Modeling and control of
zero-sequence current in multiple grid connected converter”, Proc. IEEE Power Electronics
Specialists Conference (PESC 2008), Rhodes, pp. 2064-9.
Hyunbae, K. and Lorenz, R.D. (2004), “Carrier signal injection based sensorless control methods
for IPM synchronous machine drives”, Proc. 39th Annual Meeting of IEEE Industry
Applications, Vol. 2, pp. 977-84.
Inoue, Y., Morimoto, S. and Sanada, M. (2008), “Output maximization using direct torque control High-power
for sensorless variable wind generation system employing IPMSG”, Proc. 13th Power
Electronics and Motion Control Conference (EPE-PEMC 2008), pp. 1859-65. wind energy
Jang, J.H., Sul, S.K., Ha, J.I., Kozo, I. and Mitsujiro, S. (2003), “Sensorless drive of surface-mounted generation
permanent-magnet motor by high-frequency signal injection based on magnetic saliency”,
IEEE Trans. on Industry Applications, Vol. 39 No. 4, pp. 1031-9.
Johnson, J.P., Ehsani, M. and Guzelgunler, Y. (1999), “Review of sensorless methods for 67
brushless DC”, Proc. Thirty-Fourth Annual Meeting of IEEE Industry Applications, Vol. 1,
pp. 143-50.
Kato, S., Michihira, M. and Tsuyoshi, A. (2006), “Modeling and simulation of a permanent magnet
synchronous machine with six-phase stator winding for renewable energy applications”, Proc.
Int. Conf. Electrical Machines Systems (ICEMS), Nagasaki, CD-ROM, Paper 204.
Kawabe, I., Morimoto, S. and Sanada, M. (2007), “Output maximization of wind generation
system using sensorless controlled IPMSG”, Proc. 2007 European Conference on Power
Electronics and Applications, pp. 1-10.
Khan, M., Saleh, S.A. and Rahman, M.A. (2009), “Generation and harmonics in interior
permanent magnet wind generator”, Proc. IEEE International Electric Machines and
Drives Conference, 2009 (IEMDC 2009), pp. 17-23.
Kim, H.W., Kim, S.S. and Ko, H.S. (2010), “Modeling and control of PMSG-based variable-speed
wind turbine”, Electric Power Systems Research, Vol. 80 No. 1, pp. 46-52.
Koonlaboon, S. and Sangwongwanich, S. (2005), “Sensorless control of interior
permanent-magnet synchronous motors based on a fictitious permanent-magnet
flux model”, Proc. Fourtieth Annual Meeting of IEEE Industry Applications, Vol. 1,
pp. 311-18.
Krug, D., Bernet, S., Fazel, S.S., Jalili, K. and Malinowski, M. (2007), “Comparison of 2.3-kV
medium-voltage multilevel converters for industrial medium-voltage drives”, IEEE Trans.
on Industrial Electronics, Vol. 54 No. 6, pp. 2979-92.
Larrinaga, S.A., Rodriguez-Vidal, M.A., Oyarbide, E. and Apraiz, J.R.T. (2007), “Predictive
control strategy for DC/AC converters based on direct power control”, IEEE Trans. on
Industrial Electronics, Vol. 54 No. 3, pp. 1261-71.
Li, J.L., Hu, S.J., Li, M., Zhu, Y., Kong, D.G. and Xu, H.H. (2008), “Research on the application of
parallel back-to-back PWM converter on direct-drive wind power system”, Proc. Third
International Conference on Electric Utility Deregulation and Restructuring and Power
Technologies (DRPT 2008), Nanjing, pp. 2504-8.
Li, S., Haskew, T.A. and Xu, L. (2010), “Conventional and novel control designs for
direct driven PMSG wind turbines”, Electric Power Systems Research, Vol. 80 No. 3,
pp. 328-38.
Li, Y., Zhu, Z.Q., Howe, D. and Bingham, C.M. (2007), “Improved rotor position
estimation in extended back-EMF based sensorless PM brushless AC drives with
magnetic saliency”, Proc. IEEE International Electric Machines & Drives Conference
(IEMDC 2007), pp. 214-19.
Li, Y., Zhu, Z.Q., Howe, D., Bingham, C.M. and Stone, D.A. (2009), “Improved rotor-position
estimation by signal injection in brushless AC motors, accounting for
cross-coupling magnetic saturation”, IEEE Trans. on Industry Applications, Vol. 45
No. 5, pp. 1843-50.
Malinowski, M. (2001), “Sensorless control strategies for three-phase PWM rectifiers”, PhD thesis,
Warsaw University of Technology, Warsaw.
COMPEL Malinowski, M., Jasinski, M. and Kazmierkowski, M.P. (2004), “Simple direct power control of
three-phase PWM rectifier using space-vector modulation (DPC-SVM)”, IEEE Trans. on
32,1 Industrial Electronics, Vol. 51 No. 2, pp. 447-57.
Marchesoni, M., Mazzucchelli, M. and Tenconi, S. (1988), “A non conventional power converter
for plasma stabilization”, Proc. Power Electron. Spec.Conf., pp. 122-9.
Matsui, K., Murai, Y., Watanabe, M., Kaneko, M. and Ueda, F. (1993), “A pulsewidth-modulated
68 inverter with parallel connected transistors using current-sharing reactors”, IEEE Trans.
on Power Electronics, Vol. 8 No. 2, pp. 186-91.
Meynard, T.A. and Foch, H. (1992), “Multi-level conversion: high voltage choppers and
voltage-source inverters”, Proc. 23rd Annual IEEE Power Electronics Specialists
Conference (PESC1992), Vol. 1, pp. 397-403.
Michalke, G. and Hansen, A.D. (2010), “Modelling and control of variable speed wind turbines for
power system studies”, Wind Energy, Vol. 13 No. 4, pp. 307-22.
Michalke, G., Hansen, A. and Hartkopf, T. (2007), “Control strategy of a variable speed wind
turbine with multipole permanent magnet synchronous generator”, paper presented at the
European Wind Energy Conference (EWEC2007), Milan.
Miliani, E., Ayad, M.Y., Depernet, D. and Kauffmann, J.M. (2007), “Experimental analysis of a six
phase permanent magnet synchronous generator in a variable speed constant frequency
generating system”, Proc. IEEE Twenty Second Annual Applied Power Electronics
Conference, APEC 2007, pp. 1727-32.
Morimoto, S., Sanada, M. and Takeda, Y. (2006), “Mechanical sensorless drives of IPMSM with
online parameter identification”, IEEE Trans. on Industry Applications, Vol. 42 No. 5,
pp. 1241-18.
Morimoto, S., Kawamoto, K., Sanada, M. and Takeda, Y. (2002), “Sensorless control strategy for
salient-pole PMSM based on extended EMF in rotating reference frame”, IEEE Trans. on
Industry Applications, Vol. 38 No. 4, pp. 1054-61.
Morimoto, S., Tong, Y., Takeda, Y. and Hirasa, T. (1994), “Loss minimization control of
permanent magnet synchronous motor drives”, IEEE Trans. on Industrial Electronics,
Vol. 41 No. 5, pp. 511-17.
Morren, J. and de Haan, S.W.H. (2005), “Ride through of wind turbines with doubly-fed induction
generator during a voltage dip”, IEEE Trans. on Energy Conversion, Vol. 20 No. 2,
pp. 435-41.
Muljadi, E., Yildirim, D., Batan, T. and Butterfield, C.P. (1999), “Understanding the
unbalanced-voltage problem in wind turbine generation”, Proc. IEEE Industry
Applications Conference, Vol. 2, pp. 1359-65.
Muller, S., Deicke, M. and De Doncker, R.W. (2002), “Doubly fed induction generator systems for
wind turbines”, IEEE Ind. Appl. Mag., Vol. 8 No. 3, pp. 26-33.
Nabae, A., Takahashi, I. and Akagi, H. (1981), “A new neutral-point-clamped PWM inverter”,
IEEE Trans. on Industry Applications, Vol. IA-17, p. 518.
Nakamura, T., Morimoto, S., Sanada, M. and Takeda, Y. (2002), “Optimum control of IPMSG for
wind generation system”, Proc. Power Conversion Conference, 2002 (PCC Osaka 2002),
Vol. 3, pp. 1435-40.
Nass, I., Undeland, B. and Gjengedal, T. (2002), “Methods for reduction of voltage unbalance in
weak grids connected to wind plants”, paper presented at IEEE Workshop on Wind Power
and the Impacts on Power Systems, Norway.
National Grid Transco (2004), “Appendix 1: extracts from the grid code-connection conditions”,
available at: www.nationalgrid.com (accessed 30 September 2010).
Ogasawara, S., Takagaki, J., Akagi, H. and Nabae, A. (1992), “A novel control scheme of a parallel High-power
current-controlled PWM inverter”, IEEE Trans. on Industry Applications, Vol. 28 No. 5,
pp. 1023-30. wind energy
Paicu, M.C., Boldea, I., Andreescu, G.D. and Blaabjerg, F. (2009), “Very low speed performance of generation
active flux based sensorless control: interior permanent magnet synchronous motor vector
control versus direct torque and flux control”, IET Electr. Power Appl., Vol. 3 No. 6,
pp. 551-61.
69
Pena, R., Clare, J.C. and Asher, G.M. (1996), “Doubly fed induction generator using back-to-back
PWM converter and its application to variable-speed wind – energy generation”, Proc. IEE
EPA, Vol. 143 No. 3, pp. 231-41.
Pena, R., Cardenas, R., Reyes, E., Clare, J. and Wheeler, P. (2009), “A topology for multiple
generation system with doubly fed induction machines and indirect matrix converter”,
IEEE Trans. on Industrial Electronics, Vol. 56 No. 10, pp. 4181-93.
Petersson, A. (2005), “Analysis, modeling and control of doubly-fed induction generators for
wind turbines”, PhD thesis, Chalmers University of Technology, Goteborg.
Petersson, A., Harnefors, L. and Thiringer, T. (2005), “Evaluation of current control methods for
wind turbines using doubly-fed induction machines”, IEEE Trans. on Power Electronics,
Vol. 20 No. 1, pp. 227-35.
Phillips, K.P. (1972), “Current-source converter for AC motor drives”, IEEE Trans. on Industry
Applications, Vol. IA-8 No. 6, pp. 679-83.
Podlesak, T.F., Katsis, D.C., Wheeler, P.W., Clare, J., Empringham, C.L. and Bland, M. (2005),
“A 150-kVA vector-controlled matrix converter induction motor drive”, IEEE Trans. on
Industry Applications, Vol. 41 No. 3, pp. 841-7.
Qiu, Z. and Chen, G. (2007), “Study and design of grid connected inverter for 2 MW wind turbine”,
Proc. 42nd IAS Annual Meeting, Conference Record of IEEE Industry Applications
Conference (IAS2007), New Orleans, LA, pp. 165-70.
Ramtharan, G., Jenkins, N. and Anaya-Lara, O. (2007), “Modelling and control of synchronous
generators for wide-range variable-speed wind turbines”, Wind Energy, Vol. 10 No. 3,
pp. 231-46.
Reyes, E., Pena, R., Cardenas, R., Clare, J. and Wheeler, P. (2008), “Control of a doubly-fed
induction generator via a direct two-stage power converter”, Proc. 4th IET Conference on
Power Electronics, Machines and Drives, 2008 (PEMD 2008), pp. 280-5.
Ribeiro, L.A.S., Degner, M.W., Briz, F. and Lorenz, R.D. (1998), “Comparison of
carrier signal voltage and current injection for the estimation of flux angle or rotor
position”, Proc. Thirty-Third Annual Meeting of IEEE Industry Applications, Vol. 1,
pp. 452-9.
Rodriguez, J., Bernet, S., Steimer, P.K. and Lizama, I.E. (2010), “A survey on
neutral-point-clamped inverters”, IEEE Trans. on Industrial Electronics, Vol. 57 No. 7,
pp. 2219-30.
Saylors, S. (2008), “Meeting North American grid codes”, paper presented at North East
Region System Operators Wind Integration Seminar, Vestas Americas, available at: www.
weican.ca/documents/2008/08021216Steve%20SaylorsTechnology%20Drivers%20%
20NE%20Region%20Wind%20Integration%20Seminar.pdf (accessed 30 September
2010).
Song, S., Kang, S. and Hahm, N. (2003), “Implementation and control of grid connected
AC-DC-AC power converter for variable speed wind energy conversion system”,
Proc. Eighteenth Annual IEEE Applied Power Electronics Conference and Exposition
(APEC 2003), Miami, FL, Vol. 1, pp. 154-8.
COMPEL Tang, F., Jin, X., Tong, Y., Liu, J., Zhou, F. and Ma, L. (2009), “Parallel interleaved grid-connected
converters in MW-level wind power generation”, Proc. IEEE International Electric
32,1 Machines and Drives Conference (IEMDC 2009), Miami, FL, pp. 789-96.
Vazquez, S., Sanchez, J.A., Carrasco, J.M., Leon, J.I. and Galvan, E. (2008), “A model-based direct
power control for three-phase power converters”, IEEE Trans. on Industrial Electronics,
Vol. 55 No. 4, pp. 1647-57.
70 Vizireanu, D., Kestelyn, X., Brisset, S., Brochet, P. and Semail, E. (2006), “Experimental tests on a
9-phase direct-drive PM axial-flux synchronous generator”, Proc. Int. Conf. Electrical
Machines (ICEM), Chania, Paper PMM1-15, CD-ROM.
Wen, C., Zhu, X., Li, J. and Xu, H. (2009), “Interleaved modulation of parallel three-phase PWM
converters used for wind power system”, Proc. International Conference on Sustainable
Power Generation and Supply (SUPERGEN 2009), Nanjing, pp. 1-5.
Wheeler, P., Rodriguez, J., Clare, J.C., Empringham, L. and Weinstein, A. (2002), “Matrix
converters: a technology review”, IEEE Trans. on Industrial Electronics, Vol. 49 No. 2,
pp. 276-88.
Wu, B., Rizzo, S., Zargari, N. and Xiao, Y. (2001), “An integrated DC link choke for elimination of
motor common-mode voltage in medium voltage drives”, Proc. Conf. Rec. IEEE IAS Annu.
Meeting, Chicago, IL, Vol. 3, pp. 2022-7.
Wu, B., Pontt, J., Rodriguez, J., Bernet, S. and Kouro, S. (2008), “Current-source converter and
cycloconverter topologies for industrial medium-voltage drives”, IEEE Trans. on
Industrial Electronics, Vol. 55 No. 7, pp. 2786-97.
Xu, L. and Cartwright, P. (2006), “Direct active and reactive power control of DFIG for wind energy
generation”, IEEE Trans on. Energy Conversion, Vol. 21 No. 3, pp. 750-8.
Xu, L. and Wang, Y. (2007), “Dynamic modeling and control of DFIG based wind turbines under
unbalanced network conditions”, IEEE Trans. on Power System, Vol. 22 No. 1, pp. 314-23.
Xu, Z. and Rahman, M.F. (2007a), “An adaptive sliding stator flux observer for a
direct-torque-controlled IPM synchronous motor drive”, IEEE Trans. on Industrial
Electronics, Vol. 54 No. 5, pp. 2398-406.
Xu, Z. and Rahman, M.F. (2007b), “Direct torque and flux regulation of an IPM synchronous
motor drive using variable structure control approach”, IEEE Trans. on Power Electronics,
Vol. 22 No. 6, pp. 2487-98.
Xu, Z., Ge, P. and Xu, D. (2009), “High performance control of a permanent magnet wind power
generation system using an adaptive sliding observer”, Proc. International Conference on
Power Electronics and Drive Systems (PEDS 2009), pp. 67-71.
Ye, Z., Boroyevich, D., Jae-Young, C. and Lee, F.C. (2002), “Control of circulating current in two
parallel three-phase boost rectifiers”, IEEE Trans. on Power Electronics, Vol. 17 No. 5,
pp. 609-15.
Zhang, L., Watthanasarn, C. and Shepherd, W. (1997), “Application of a matrix converter for the
power control of a variable-speed wind-turbine driving a doubly-fed induction generator”,
Proc. 23rd International Conference on Industrial Electronics, Control and
Instrumentation, New Orleans, LA, Vol. 2, pp. 906-11.
Zheng, L., Fletcher, J.E., Williams, B.W. and He, X. (2008), “Dual-plane vector control of a
five-phase induction machine for an improved flux pattern”, IEEE Trans. on Industrial
Electronics, Vol. 55 No. 5, pp. 1996-2005.
Zhi, D. and Xu, L. (2007), “Direct power control of DFIG with constant switching frequency and
improved transient performance”, IEEE Trans. on Energy Conversion, Vol. 22 No. 1,
pp. 110-18.
About the authors High-power
Z.Q. Zhu received the BEng and MSc degrees in Electrical and Electronic wind energy
Engineering from Zhejiang University, Hangzhou, China, in 1982 and 1984,
respectively, and the PhD degree in Electrical and Electronic Engineering from
generation
the University of Sheffield, Sheffield, UK, in 1991. From 1984 to 1988, he was a
Lecturer with the Department of Electrical Engineering, Zhejiang University.
Since 1988, he has been with the University of Sheffield, where he was initially a 71
Research Associate and was subsequently appointed to an established post as
Senior Research Officer/Senior Research Scientist. Since 2000, he has been a Professor of
Electrical Machines and Control Systems with the Department of Electronic and Electrical
Engineering, University of Sheffield, and is currently Head of the Electrical Machines and Drives
Research Group. His current major research interests include design and control of permanent
magnet brushless machines and drives, for applications ranging from automotive and aerospace
to renewable energy. He is a Fellow of IEEE. Z.Q. Zhu is the corresponding author and can be
contacted at: Z.Q. Zhu@sheffield.ac.uk

Jiabing Hu received the BSc and PhD degrees from the College of Electrical
Engineering, Zhejiang University, Hangzhou, P.R. China, in July 2004 and
September 2009, respectively. From 2007 to 2008, he was as a Visiting Scholar
with the Department of Electronic and Electrical Engineering, University of
Strathclyde, Glasgow, UK. He is currently a Post-Doctoral Research Associate
with the Department of Electronic and Electrical Engineering, University of
Sheffield, Sheffield, UK. His current research interests include motor drives,
digital control and application of power electronics in renewable energy conversion, especially
the control and operation of doubly fed induction generator (DFIG) and permanent magnet
synchronous generator (PMSG) for wind power generation. He is a Member of IEEE.

To purchase reprints of this article please e-mail: reprints@emeraldinsight.com


Or visit our web site for further details: www.emeraldinsight.com/reprints

You might also like