You are on page 1of 48

18

CHAPTER 2

PERMANENT MAGNET SYNCHRONOUS GENERATOR

2.1 GENERAL

Variable speed wind turbines are increasingly being used for


electrical power generation as they capture more energy than the fixed speed
wind turbines. The variable speed operation of a wind turbine can be achieved
with a Doubly Fed Induction Generator (DFIG) or with a Permanent Magnet
Synchronous Generator (PMSG). The PMSG has many advantages over the
DFIG. The PMSG does not require DC excitation as the magnetic field is
produced by the permanent magnets rather than by the coil. Hence, the PMSG
does not require slip rings and brushes, which reduces the weight, cost, losses,
and maintenance(Samuel et al 2010). However, large permanent magnets are
costly, which restricts the economic rating of the machine. Variable-speed
wind power generation with PMSG and a full-scale power electronic
converter is a promising but not yet a very popular wind turbine concept. A
PMSG, connected to the power electronic converter, can operate at low
speeds; hence, a gearbox is not required. Since the gearbox increases the
weight, the losses, the cost, and demands more maintenance, a gearless
construction represents an efficient and robust solution, especially for
offshore applications. PMSGs are classified as radial flux PMSGs, axial flux
PMSGs and transverse flux PMSGs based on the direction of the flux lines.
They are also classified as surface inset PMSGs, surface mounted PMSGs and
interior PMSGs based on the location of the permanent magnets on the rotor
(Boldea 2005). The surface mounted PMSGs are the most efficient (Sebastian
19

& Slemon 1989). Figure 2.1 shows the cross-sectional view of a four-pole
surface mounted PMSG.

Stator Core

Permanent
Magnets

Rotor Core

Figure 2.1 Structure of the permanent magnet synchronous machine

The stator carries a three-phase winding, in which the three-phase


emfs are induced (Belakehal et al 2009). The magnets are mounted on the
surface of the rotor core and have the same role as the field windings in a
synchronous machine except that their magnetic field is constant and there is
no control over it. When the rotor is rotated by external means, the magnetic
field of the permanent magnet cuts the three-phase windings, resulting in
induced emfs on the three-phase stator windings.

2.2 LITERATURE REVIEW

Samuel et al (2010) described a cascaded H-bridge multilevel


inverter (CHB-MLI) based wind energy conversion system (WECS) topology
with the permanent magnet synchronous generator (PMSG). They drove the
PMSG by a variable speed wind turbine, which feeds a high power single-
phase load and connected the output of the PMSG to a three-phase
20

uncontrolled rectifier whose output is connected to the load through the H-


bridge inverter.

Sebastian & Slemon (1989) modeled the surface-mounted


permanent magnet motor in the d-q reference frame. They validated the
circuit model by conducting tests on the motor with neodymium-iron-boron
magnets, built in a 5 HP induction motor frame, and showed that there exists
negligible effect of the magnetic conductivity on the dynamic response. They
developed an expression for the maximum stator current, which can be used
without magnetic damage. In addition, they showed experimentally that a
linear relationship exists between the torque and the stator current. The linear
relation exists until the stator current is six times the rated value, and the
torque is about nine times its rated value.

Belakehal et al (2009) described an isolated PMSG based wind


energy conversion system. They connected the system to a DC load through a
diode rectifier, a boost converter and a battery. They modeled the wind
turbine and the PMSG and explained the aerodynamic characteristics of the
wind turbine and the power conversion system topology. Furthermore, they
developed the maximum power tracking of the wind turbine generator system
using MATLAB and showed that the developed maximum power-tracking
algorithm is suitable for the wind turbine generation system.

Hemeida et al (2011) analyzed the vector control method for an


ultra large wind turbine by developing a Fatigue, Aerodynamics, Structures
and Turbulence (FAST) nonlinear wind turbine model in Simulink. The
FAST model allows the control of the generator torque, the pitch angle, the
yaw angle of the nacelle, and the high-speed shaft (HSS) brake. TurbSim,
which is a full field wind simulator developed by the National Renewable
Energy Laboratory (NREL) provides the input to the FAST model. They
modeled the multi-pole radial flux PMSG and proved it a good choice for
21

wind turbines. In addition, they controlled the system by the generator side
converter controller and the grid side converter controller and studied the
performance of the controllers, namely the vector controller and the pitch
controller.

Bao et al (2012) proposed a WECS with a Multilevel Current


Source Inverter (MCSI). They derived the MCSI from the Multilevel Voltage
Source Inverter (MVSI) so that the existing modulation and control schemes
of the MVSI could be used. In addition, they controlled the DC-link current
by a DC-link current controller and controlled the real power and the reactive
power in the grid independently by the MCSI.

Kumar & Hasnat (2014) solved problems of harmonic currents and


negative sequence currents introduced by micro grids by means of multilevel
inverters with controllers. They used high voltage and high power three-level
neutral point clamped converter for the PMSG, which decreases the output
harmonics, improves the power capacity, and reduces the voltage stress on the
switches. Moreover, they achieved the maximum power tracking at the
generator side converter, which also operates the generator in a stable and
efficient manner with the maximum torque to current ratio. They also
achieved decoupled control of the real and reactive powers by the grid side
controller through the voltage-oriented control.

Samuel et al (2011) proposed a WECS with the split winding


alternator based on a cascaded H-bridge multilevel inverter for the grid side
converter. They operated the inverter in the current control mode to achieve
power factor correction, harmonic compensation and load balancing. In
addition, they proposed two controls (i) slow DC-link voltage control by
controlling the set terminal voltage of the alternator by the excitation
controller, (ii) instantaneous current control to achieve power factor
correction, harmonic compensation and load balancing. They also proposed a
22

new method of generating reference currents for the cascaded H-bridge


multilevel inverter.

Wu et al (2013) proposed a new direct driven PMSG wind energy


system containing an auxiliary converter connected in parallel with the grid
side converter and a coordinated control strategy to enhance and improve the
low voltage ride-through (LVRT) capability and the power quality of the
system. During normal operation, the main grid-side converter maintains the
DC-link voltage constant, whereas the auxiliary grid-side converter functions
as an active power filter (APF). The APF improves the power quality with
harmonic suppression and reactive power compensation. During grid faults, a
hierarchical coordinated control scheme for the generator-side converter, main
grid-side converter and auxiliary grid-side converter is presented, depending
on the grid voltage sags, to enhance the LVRT capability of the direct-driven
PMSG wind turbine. John et al (2013) implemented an asymmetric cascaded
H-bridge multilevel inverter for the grid connected wind energy conversion
system with the PMSG, to suppress the total harmonic distortion (THD).

1 Wilamowski & Irwin (2011) dealt with the construction of the


PMSG and their types. He modeled the machine and presented the vector
control. Barnes (2003) presented the sensor-less vector control of variable
speed AC drives. Krishnan (2010) dealt with the dynamic and state space
models of the permanent magnet synchronous machine (PMSM) and the
various vector control strategies.

Yuan et al (2012) proposed a modular high power, medium voltage


converter for a wind energy conversion system with large PMSG that is free
from the grid side transformers. They cascaded the converters to obtain
medium voltage, the converters being fed from a pair of generator coils,
which are 90º phase shifted to achieve a stable DC-link power. The generator
side converter is modeled and controlled with two controllers, the outer DC-
23

link voltage control loop and the inner current control loop. The outer loop
maintains the DC-link voltage of the converter under different wind speeds
and provides the reference current amplitude for the inner current loop. The
inner current loop enables the coil current to be sinusoidal and track the
generator back-EMF. The grid side control is achieved by the cascaded H-
bridge converter through vector control.

Yaramasu et al (2013) proposed a configuration consisting of a


medium-voltage PMSG connected to a three-phase diode bridge rectifier, a
DC-DC four-level boost converter as the intermediate stage, and a four-level
diode-clamped inverter on the grid-side. The DC-link capacitor voltages are
balanced by the boost converter, which simplifies complexity of the grid-tied
inverter. They developed a cost function, which includes the DC-link voltage
regulation, reactive power generation to meet the grid requirement, balancing
of capacitor DC-link voltage and minimization of switching frequency. In
addition, they proposed a voltage-oriented Finite-Control Set Model-
Predictive Control (FCS-MPC) for the grid tie inverter.

Chen et al (2009) reviewed the power electronic applications for


wind energy conversion system. They reviewed the various wind turbine
systems with various wind generators and power electronic converters. In
addition, they studied the different methods of maximum power point
tracking (MPPT). They also reviewed the power quality issues such as the
voltage control, reactive power control, voltage stability, and active power
handling.

Xia et al (2010) studied a direct driven wind energy conversion


system with a three-level boost converter and a three-level neutral point
clamped multilevel inverter. They achieved the neutral point potential
balancing by a new method called the switch-signal phase delay control
(SSPDC). They also achieved the maximum power point tracking with the
24

boost chopper and the two PI regulators. They compared the effects of
reactive power control by the SSPDC and hysteresis control methods and
showed that the range of capacity for the reactive power compensation under
the SSPDC is wider than that under the hysteresis control.

Yaramasu et al (2011) proposed a new topology using diode


rectifier, three-level boost (TLB) and neutral point clamped (NPC) converters
to reduce the cost and size. They proposed a DC-link maximum power point
tracking (MPPT) control scheme with which the TLB converter performs the
MPPT and balancing of DC-link capacitors that provides a greater flexibility
for NPC control.

Li et al (2010) developed conventional and novel control methods


for the direct driven PMSG based WECS. They modeled the PMSG in the d-q
reference frame for the transient as well as the steady state condition. In
addition, they proposed a novel vector control technique integrating the PID,
fuzzy and adaptive control techniques in an optimal control configuration.
The proposed control method is found to be superior in achieving the desired
control objectives of the PMSG and the optimal performance of the whole
system.

Rolan et al (2009) modeled the PMSG in the synchronous reference


frame. They also modeled the wind speed, wind turbine and the drive train
and presented the MPPT concept in terms of generator rotor speed and wind
speed. The models are implemented in the MATLAB/Simulink.

Mittal et al (2009) proposed and modeled a variable speed PMSG


based WECS. They discussed the various ride-through topologies of the grid
connected PMSG based wind energy system. They also discussed the grid
interconnection issues during healthy and faulty conditions and studied the
effects of symmetrical and unsymmetrical faults.
25

Sasikumar & Chenthur Pandian (2011) considered a variable speed


WECS with the Self Excited Induction Generator (SEIG). The system uses a
cascaded H-bridge multilevel inverter at the load side to reduce the total
harmonic distortion (THD). The SEIG is modeled in the synchronously
rotating d-q reference frame. They simulated the entire system with the SEIG,
diode rectifier, multilevel inverter, and the load in MATLAB for the RL and
the induction motor loads. The THD at the load side was minimised.

In the literature, Dai et al (2009) have proposed a PMSG based


WECS with a current source inverter for the grid connection. They performed
the generator side control based on the synchronously rotating reference
frame with rotor flux orientation. They also proposed the grid side control
using the power feed-forward method for a stable DC-link current operation.

Margaris et al (2011) investigated dynamic security issues in


isolated power systems based on the detailed models of the power system, and
of the most common wind turbine configurations together with their fault
ride-through controllers. They also investigated the power system issues
during their connection to the grid with the detailed generic models for three
wind turbine technologies such as Active Stall Induction Generator (ASIG),
Doubly Fed Asynchronous Generator (DFAG) and PMSG. They showed that,
next to the protection of the non-interconnected power systems, the Fault Ride
through (FRT) capability and the response of the wind farms during and after
disturbances determine the load shedding and the frequency deviations. The
FRT capability improves the system response during faults ensuring limited
load shedding and frequency dips.

Boldea (2005) described and modeled all types of variable speed


generators. He also presented the performance and control of variable speed
generators. Sanchez et al (2008) presented a dynamic model for a variable
speed WECS with variable pitch, synchronous electrical generator and full
26

scale back-back power converters. They showed that the increase in


computational time is 66% with the developed model.

Saikumar & Sivakumar (2011), Maiti et al (2008), Kim et al (1995)


proposed a model reference adaptive control based speed controller for the
permanent magnet synchronous motor. Two models are needed for the
controller; hence, the complexity of the system increases. Zhang et al (2009)
compared the sensor-less control of the induction machine with the
Luenberger observer, sliding mode observer and the extended Kalman filter.
They proved that the first two methods are more suitable than the third
method.

Khan & Iravani (2008) proposed a hybrid valve switching and


control strategy for grid connected voltage source converter of a WECS.
They showed that with the proposed hybrid switching control and a simple
linear regulator, the grid connected WECS could have the ride-through
capability for single-phase temporary faults. In addition, they showed that the
system operates without any intermediate energy storage during the transient
and steady-state conditions.

Fan et al (2010) implemented the sensor-less vector control of


speed and position with a digital phase-locked loop. They controlled the
output frequency and measured the motor speed and the rotor position angle
from the obtained frequency characteristic. However, this frequency control
technique does not work well when the motor runs at low speed. Zhonggang
et al (2012) developed a sensor-less vector control with a Robust Extended
Kalman Filter (REKF) algorithm. They also analysed the Extended Kalman
Filter (EKF) algorithm. The speed estimation error, the flux change due to
external gross error and the internal estimation gross error based on REKF are
compared with that of EKF. They showed that the REKF gave better
performance than EKF.
27

2 Zambra et al (2010) compared the three topologies of multilevel


inverters namely, the neutral point clamped (NPC), symmetrical cascaded and
hybrid asymmetrical cascaded multilevel inverters. They proved that the
cascaded topology with a single DC source and transformers is better than the
NPC even in terms of cost and distortion in the output voltage. Kangarlu &
Babaei (2012) proposed a generalized multilevel inverter with reduced
number of switches and analyzed it for symmetric and asymmetric conditions.

Ma & Blaabjerg (2011) designed and compared a 10 MW PMSG


based wind generation system with various multi-level converter
configurations. They showed that the thermal performance of the power
devices decides the life of the converters. They also showed that the thermal
performance of the CHB converter is better when compared to that of the
NPC topology.

3 Rech et al (2002) analyzed different hybrid multilevel inverters for


high-power applications that minimize the number of series-
connected inverters for the given level. They compared the number of levels,
power, and spectral performances for various configurations of the
hybrid multilevel inverters. They also proposed a design procedure to obtain
the number of H-Bridge cells to produce an output voltage with more number
of levels with fewer H-Bridge cells, in order to reduce the circulating energy
between the cells.

4 Rasheed et al (2013) analyzed two types of three-phase inverters


such as the CHB and the Diode Clamped inverters and stated that the THD of
the output for both the inverters decreases. Bahravar et al (2012) proposed a
new basic topology comprising of two DC voltage sources, two unidirectional
power switches, and an H-bridge for the cascaded multilevel inverters. The
proposed topology requires less number of switches for the symmetric
cascaded inverters.
28

5 Peng (2001) proposed a generalized multilevel inverter topology


with self-voltage balancing. The generalized topology balances each DC
voltage level automatically without any assistance from the other circuits.
Using this topology, any existing and new structures of multilevel inverters
can be obtained. Floricau (2013) proposed two new DC/AC
hybrid multilevel bidirectional switching cells, such as the three-voltage level
cells and the five-voltage level cells. New single-phase and three-phase
multilevel inverters can be developed using the proposed switching cells.

Laali et al (2010) proposed and analyzed various modulation


techniques of the multilevel inverter and presented the multicarrier
modulation techniques. Chinnaya et al (2007) simulated the cascaded H-
bridge inverters with the pure sine pulse width modulation (SPWM) and the
third harmonic injection PWM, for varying modulation indices. Though the
third harmonic injection PWM increases the gain of the inverter, the presence
of third harmonic causes problem when the load neutral is grounded. Calais et
al (2001), Urmila & Subbarayudu (2010), Colak et al (2011), McGrath &
Holmes (2002) discussed the various multicarrier PWM techniques. In the
carrier disposition techniques, the carrier waves are vertically shifted from
each other. In the phase-shifted technique, the carrier waves have a horizontal
phase shift.

Liserre et al (2011) reviewed the various wind generators, power


electronic converters, generator side control and grid side issues. They studied
the doubly fed induction generator, the direct drive synchronous generator,
and the PMSG with the back-back converters. They also dealt with the
two-level PWM inverters, multilevel inverters and the matrix converters.
29

Blaabjerg et al (2004) dealt with modern power electronic


development of distributed generation systems like wind turbine generation
system, photovoltaic (PV) system and fuel cells. Various power electronic
converters used for the distributed generation system were discussed. They
concluded that the power electronic converters play a vital role in the systems.
Li & Chen (2008) compared the different wind generation systems, and
investigated the various types of PMSGs. They showed that the variable speed
wind energy system with power electronic converters is the best choice for
large wind farms.

Blaabjerg et al (2006), Altin et al (2010), Tsili & Papathanassiou


(2009) described the back-to-back power electronic converters for the wind
turbines, which enable the full control of the extracted power. Power
electronic converters improve the performance of the wind energy system by
increasing the energy yield and reducing the mechanical stress and play a vital
role in maintaining the quality of the power injected into the grid.

Blaabjerg & Ma (2013) discussed various developments in


technologies for the wind energy system and reviewed the power electronic
converters, control systems, and grid connection issues. They addressed the
future technologies for wind generation system.

Denniston et al (2011) described a new approach for high-gain,


high-voltage DC-DC converters using multiple modules of single-switch
single-inductor transformer-less converters. In offshore wind turbines,
multiple module converters yield high gains and efficiencies due to the
absence of transformers. They compared the conventional converters with the
multiple module converters and found the latter to be superior in terms of
device rating, device count and reliability.
30

6 Zhao & Lee (2003) presented simple topologies for a family of


high-efficiency, high step-up, clamp-mode coupled-inductor DC-DC converters
with theoretical analysis, practical design and experimental results. They
replaced the active switches with one additional diode and one
coupled winding to perform the function of active clamps. The leakage energy
is recovered by adding a small clamp capacitor and the reverse-recovery
problem of the output rectifier is alleviated by using the leakage inductor.
Hence, high efficiency is achieved.

7 Senjyu et al (2006) and Muyeen et al (2007) proposed the pitch


angle control for a wind turbine to maintain the wind power at a constant
value to avoid the wind power fluctuation due to the randomly varying wind
speed. Cardenas et al (2004) proposed a flywheel energy system driven by
vector controlled induction machine to smooth out the fluctuations in wind
power. Nomura et al (2005) and Ali et al (2006) have used a superconducting
magnetic energy storage system to maintain the wind power at a constant
value. Muyeen et al (2009) proposed an energy capacitor system for a PMSG
based variable-speed wind turbine to smoothen the wind power.

8 Bhende et al (2011) presented an algorithm for the effective


energy management of a standalone variable speed WECS with the PMSG.
The DC-link voltage is maintained constant irrespective of the variations in
the wind speed and the load to maintain the inverter voltage a constant value.
The line voltages at the point of common coupling (PCC) are maintained
balanced, even with the unbalanced load by an effective pulse width
modulation control technique of the inverter. The DC-DC converter controller
in addition to maintaining the constant DC voltage also act as a DC side
active filter and reduces the oscillations of the generator torque caused by the
unbalanced load.
31

9 Rashid (2007), Kassakian et al (2010), Sen (2010), Singh (2007)


analyzed the buck, boost, buck-boost, and Cuk converters. The waveforms of
the inductor current, supply current and capacitor voltage are presented. The
differential equations during the ON period and the OFF period of the switch
and the average output voltage and design equations are derived. The
expressions for the ripple current and voltage are also derived. Bose (2002)
dealt with the single-phase and three-phase diode rectifiers. The expressions
for the average output voltage is also derived.

Tao et al (2008) described a three-port bidirectional DC-DC


converter with zero-voltage switching. They achieved soft-switching
conditions over the full operating range by changing the duty ratio of the
voltage applied to the transformer winding. A dual PI based control strategy is
proposed to obtain the zero-voltage switching, constant output voltage and
power flow management. Yaramasu & Wu (2011) developed a topology
using the diode rectifier, three-level boost converter and an NPC converter for
a medium voltage PMSG based WECS. The three-level boost converter
performs the maximum power point tracking and balances the DC-link
capacitors using a DC-link maximum power point tracking control scheme,
which provides greater flexibility for the NPC control.

2.3 MODELING OF THE PMSG

The mathematical model of the permanent magnet synchronous


machine is based on the assumptions that the stator windings are positioned
sinusoidal along the air-gap as far as the mutual flux effect with the rotor is
concerned; the rotor inductances do not vary much with the variation of the
rotor position with respect to the stator slots. In addition, the stator windings
are considered symmetrical, the resistances of the windings are constant, with
32

the damper windings and capacitances neglected. Hence, the power losses
remain constant, when the magnetic hysteresis and saturation effects are
neglected (Kumar & Maheshan 2012). The circuit model starts with the
phase coordinate model. Equations (2.1) - (2.3) give the voltage equations of
the stator phases (Boldea 2006)

d a
va ia Rs (2.1)
dt

d b
vb ib R s (2.2)
dt

d c
vc ic R s (2.3)
dt

where, va, vb, and vc are the phase voltages of the stator phases ‘a’, ‘b’, and ‘c’
respectively; ia, ib, and ic are the stator currents of the phases ‘a’, ‘b’, and ‘c’
respectively; and a, b, and c are the flux linkages of the phases ‘a’, ‘b’,
and ‘c’ respectively. Equation (2.4) gives the flux linkages in the matrix form

a L L aa L L ca ia ma er
sl er ab er er
L L L L i
b ab er sl bb er bc er b mb er
L ca L L L cc ic
c er bc er sl er mc er

(2.4)

where, Laa, Lbb, Lcc are the self inductances of the stator phases ‘a’, ‘b’, and ‘c’
respectively; Lab, Lbc, Lca are the mutual inductances between phases ‘a’ and
‘b’, ‘b’ and ‘c’ and ‘c’ and ‘a’ respectively; Lsl is the leakage inductance while
ma, mb, and mc are the permanent magnet flux linkages of the phases ‘a’,
‘b’, and ‘c’ respectively. er (rotor position) is the angle between the
permanent magnet pole axis and the ‘a’ phase.
33

2.4 THE DQ MODEL

Equations (2.5) and (2.6) give the voltage equations of permanent


magnet synchronous machine in the d-q reference frame with the d-axis
coinciding with the permanent magnet pole axis

d
v Rs i d (2.5)
ds ds e q
dt

d q
v qs Rs iqs e d (2.6)
dt

where, e is the electrical speed in rad/s, d, and q are the d-axis and the q-
axis flux linkages respectively. vds, and vqs are respectively the stator d and q
axes voltages in the synchronously rotating reference frame. ids, and iqs are
respectively the d and q axes stator currents in the synchronously rotating
reference frame. Equations (2.7) and (2.8) give the flux linkages in the d-q
reference frame

L i m (2.7)
d d ds

q Lqiqs (2.8)

where, Ld, and Lq are the d and q-axes inductances respectively, and m is the
permanent magnet flux linkage.

Equations (2.9) and (2.10) respectively give the d-axis and the q-
axis voltage equations obtained by substituting the values of d and q.

Figure 2.2 gives the d-axis and q-axis equivalent circuits.

d L i
v ds R s i ds d ds (2.9)
e L q i qs
dt
34

d L i
q qs
vqs Rs iqs e Ld ids (2.10)
dt m

Equation (2.11) gives the electromagnetic power from which the


electromagnetic torque is derived.

3
Pem e i i (2.11)
2 d qs q ds

Equation (2.12) gives the relation between the electrical speed and
the mechanical speed

e p m (2.12)

where, p is the number of pole pairs and m is the mechanical speed of the
rotor in rad/s. Hence, Equation (2.13) gives the expression for the
electromagnetic torque. By substituting the d-axis and q-axis flux linkages
in Equation (2.13), we get the torque Equation (2.14).

Pem 3
Te p i i (2.13)
m 2 d qs q ds

Rs Ld Rs Lq

ids iqs
vds vqs
e q e d

Figure 2.2 (a) d-axis equivalent circuit, (b) q-axis equivalent circuit
35

Pem 3
Te p i L L i i (2.14)
m 2 m qs d q qs ds

For the surface mounted PMSG, the d and q axes inductances are
equal and hence Equation (2.15) gives the torque equation.

3
Te p i (2.15)
2 m qs

The d and q-axes currents have negative signs for the generating
machine; hence the electromagnetic torque and power are also negative.

2.5 WIND ENERGY CONVERSION SYSTEM WITH PWM


COVERTERS

Figure 2.3 shows the WECS with the PWM back-back converters.
The wind turbine, which converts the mechanical energy of the wind turbine
into electrical energy, is connected to the PMSG. The output of the PMSG is
connected to the PWM converter, which is called the generator side converter.
The generator side converter is a voltage source converter that converts the
PMSG voltage into the DC voltage. The generator side converter controller
regulates the speed by adjusting the speed of the PMSG to match the optimum
speed of the wind turbine. The optimum speed is that speed at which the wind
turbine captures the maximum power from wind. The grid side converter
converts the DC-link voltage into an AC voltage of desired frequency. The
grid side converter controller maintains the DC-link voltage constant by
controlling the power flow. This control strategy controls the DC bus voltage
at the constant value to regulate the power delivered by the PMSG to the grid
(Emna et al 2013).
36

Generator side Grid side converter


converter
Transformer

Filter
PMSG AC/DC DC/AC
Grid

Generator Side Grid Side Converter


Converter Control Control

Figure 2.3 The WECS with the PWM back-back converters

2.5.1 The Generator Side Converter Control

There are two control techniques for the PMSG for variable speed
applications, namely the field oriented control and the direct torque control.
The field oriented control technique has less current distortion and therefore
has better performance, more overall efficiency and higher power factor when
compared to the direct torque control technique.

2.5.1.1 Generator side converter

The generator side converter (GSC) is a voltage source converter


(VSC) that acts as a rectifier. Figure 2.4 shows the circuit of the VSC.
Equation 2.16 gives the phase voltages in terms of the DC voltage and the
duty cycles (Da, Db, Dc).
37

T1 T3 T5

a
VDC
b C

T4 T6 T2

Figure 2.4 Circuit of the voltage source converter

VDC
va 2Da Db Dc
3

VDC
vb 2Db Dc Da (2.16)
3

VDC
vc 2Dc Da Db
3

Equation (2.17) gives the DC-link current in terms of the duty


cycles and the line currents (equal to phase currents of the stator).

ia
I DC Da Db Dc ib (2.17)
ic
38

2.5.1.2 Space vector modulation (SVM)

The SVM determines the duty cycles necessary for the control of
voltage, frequency and power. With SVM, three-phase quantities are
represented as vectors in a two-dimensional plane (stationary reference
frame) providing the duty cycles, which controls the power flow through the
converter. There are 8 states (6 active/non-zero states and 2 zero states)
available based on the switching states. The active states produce non-zero
voltages and the zero states produce zero voltages. The voltage of each
state depends on the space vector. Equation (2.18) gives the reference space
vector V *,

2 2
* 2 * j
*
j
V v jv va e 3
v b e 3
v c* (2.18)
3

va*, vb*, vc* are the reference phase voltages, v is real part of V *
which is
along the a-axis and v is imaginary part of V * . The active state space vectors
divide the plane into six sectors (S-I to S-VI). The reference vector V* in
any sector is obtained by switching any two adjacent vectors. Figure 2.5
shows the 8 space vectors and the six sectors. The reference space vector,
which has an angle from V1, is located in the first sector and is obtained by
switching the state vectors V1 and V2. The reference vector in sector 1 (S-I) is
bounded by four state vectors V1 (100), V2 (110), V7 (111) and V0 (000).
Figure 2.6 shows the state sequence for S-I for the three phase output
voltages, where t1, t2, t0, t7 are the time intervals of the vectors V1, V2, V0, V7
respectively. The time intervals of V0 and V7 are equal (t0= t7).
39

V3(010) V2(110)

S-II

S-I
S-III
V*
V4(011) V0(000) V1(100)
V7(111)
S-VI
S-IV

S-V

V5(001) V6(101)

Figure 2.5 The space vectors and the six sectors

000 100 110 111


va

vb

vc

t0 t1 t2 t7
Ti
Ti

Figure 2.6 The state sequence for S-I


40

Equation (2.19) gives the reference vector in sector-1,

t1 t2
V* V1 V2 (2.19)
Ti Ti

2 2 1 3
where, V1 V DC and V 2 V DC j
3 3 2 2

Also. V * V * cos j sin (2.20)

Ti is the time period of the switches of the VSC. From Equations


(2.19) and (2.20), t1 and t2 are calculated as given by Equations (2.21) and
(2.22).

V*
t1 3 Ti sin (2.21)
VDC 3

V*
t2 3 Ti sin (2.22)
VDC

From the values of t1 and t2, the value of t0 can be found. The
values of the duty cycles can be calculated using t1, t2 and t0. Equations (2.23)
- (2.25) give the expressions for duty cycles for sector-1.

t0
t1 t2
Da 2 (2.23)
Ti

t0
t2
Db 2 (2.24)
Ti
41

t0
2 t0
Dc (2.25)
Ti 2Ti

The duty cycles for other sectors are found in the same way. The
switching signals are obtained from the duty cycle values.

2.5.1.3 Field oriented control

The Field Oriented Control (FOC) is a vector control technique,


which controls the torque indirectly by controlling the stator currents. This is
achieved by using the constant torque angle control. The advantage of the
FOC is that the control is independent of the parametric variations. In the
FOC, the d-axis of the rotor is aligned with the permanent magnet axis. The
FOC uses the rotor speed as the feedback signal for its control strategy
(Hemeida et al 2011). The d-axis current component is set to zero, and
therefore the stator current has only the q-axis component. Since the torque
angle and the permanent magnet flux are constant, the torque depends only
on the q-axis stator current. Figure 2.7 represents the phasor diagram for the
FOC, with the permanent magnet flux linkage aligned along the d-axis.

Ist
q b e
d

t ids
iqs m

er a

Figure 2.7 Phasor diagram for the field oriented control


42

t is the torque angle, er is the rotor position, Ist = ids + jiqs. Figure 2.8 shows
the block diagram for the FOC with sensor, which uses one speed loop and
two current loops. The rotor speed is compared with the reference speed and
the error in speed is processed by a PI controller to provide the reference
current for the q-axis current controller. The reference value for the d-axis
current is set to zero. The rotor position is used for converting the stator
currents in the abc frame to the d-q frame. The d-q currents are compared with
the reference currents and the errors in current are processed by the current
controllers. The reference d-q voltages are generated after adding the
compensating components to the outputs of the current controllers for having
the decoupling effect. The reference d-q voltages are transformed into the
frame voltages and given as reference voltages to the SVM block (Cimpoeru
2010). The DC-link voltage is also given to the SVM block to produce the
required PWM signals for the six switches of the VSC.

IDC

GSC
VDC
PMSG (VSC)

abc

dq

PWM
er
p 1
ids iqs
s
V SVM
m
i*ds=0 PI dq
&
PWM
PI PI
*
m
i*qs
Component I
m for max
Wind speed (v ) Component II
power w

Figure 2.8 Field oriented control of the PMSG with the sensor
43

For decoupling purposes, a feed-forward loop is used that includes


components I and II, which are,

Component I: e q e iqs Lq

Component II: e d e i ds L d e m

VDC and V are given to space vector modulation block to produce the gate
signals for the converter switches. The wind speed is fed to a block where the
rotational speed of the rotor corresponding to the maximum power capture is
found. When the PMSG is operated at this speed, the power obtained from
wind has a maximum value.

2.5.1.4 q-axis current controller design

Figure 2.9 shows the block diagram for the design of the q-axis
current controller. The various blocks are the PI current controller block, the
delay introduced by the digital calculation with time constant T=1/fs, fs is
sampling frequency. (T is taken as 0.0002 s), the delay introduced by the
inverter with the time constant of 0.5Ti (Ti is taken as 0.0001 s), the plant or
the stator block, and the delay introduced by digital to analog conversion that
is taken to be equal to half of the delay introduced by the digital calculation.

i *qs 1 1 1 i
PI qs
sT 1 0.5sTi 1 sLq Rs
i Inverter
qs
Stator
Sampling
1
0.5sT 1

Figure 2.9 Block diagram for the design of the q-axis current controller
44

Lq and Rs are respectively 22 mH and 2.2 . Equation (2.26) gives the transfer
function of the q-axis current controller

1
Gq ( s ) K pq 1 (2.26)
sT
iq

where, Kpq is the proportional gain a n d Tiq is the integral time of the q-axis
current controller.

2.5.1. 5 d-axis current controller design

The d-axis current controller is implemented in the same way as


the q-axis current controller except that the Lq is replaced by Ld.
Equation (2.27) gives the transfer function of the d-axis current controller

1
G ( s)
d
K
pd
1 (2.27)
sT
id

where, Kpd is the proportional gain of the d-axis current controller and Tid is
the integral time of the d-axis current controller.

2.5.1.6 q-axis speed controller design

The various blocks include the PI speed controller, the delay


introduced by the PI controller, the delay introduced by the current controller,
the plant, and the filter with the time constant Tf and the delay introduced by
digital to analog conversion.

The current loop transfer function is derived as follows. Equation


(2.28) gives the loop transfer function of the q-axis current loop.
45

1 1 1 1 1
GH ( s) K 1 (2.28)
q pq sT sT 1 0.5sT 1 sL R 0.5sT 1
iq i q s

Since the poles introduced by the digital calculation, digital to


analog conversion and the inverter are far away, their contribution is less and
hence the loop transfer function is reduced to Equation (2.29).

1 1
GH ( s ) K 1 (2.29)
q pq sT sL R
iq q s

Equation (2.30) gives the closed loop transfer function of the q-


axis current loop. Figure 2.10 shows the block diagram for the design of the
speed controller.

1 1
K pq 1
iqs sTiq sLq Rs
(2.30)
i * qs 1 1
1 K pq 1
sTiq sLq Rs

Current loop TT
*
i qs Te
i i qs 3p m
PI 1 qs m p
sT 1 sJ B
* i *
2
m Control qs
m
algorithm
1
1
sTf 1
0.5sT 1
Filter
Sampling

Figure 2.10 Block diagram for the design of the speed controller
46

In Figure 2.10, TT is the turbine torque, m is the mechanical speed,


J is the moment of inertia and B is the viscous friction coefficient. The values
of J and B are respectively 0.0002 kgm2 and 0.0001 Nm-s/rad. Equation
(2.31) gives the transfer function of the PI speed controller,

1
G ( s) Kp 1 (2.31)
sT
i

where, K is the proportional gain of the speed controller and T is the


integral time value of the speed controller.

The PI controllers are designed using the Ziegler-Nichols second


method of tuning. The various steps involved in this method are:

1. The system is initially operated with zero integral gain


(integral time maximum).

2. The proportional gain of the controller is increased from zero


to a critical value Kcr, until a sustained oscillation is obtained.

3. The critical value of the proportional gain Kcr and the period
of the sustained oscillation Tcr are noted.

Equations (2.32) - (2.34) give the gains of the PI controller.

KP 0.45 K cr (2.32)

Tcr
Tin (2.33)
1. 2

KP
Ki (2.34)
Tin
47

Kp, Ki and Tin are respectively the proportional gain, integral gain
and integral time of PI controller. Table 2.1 shows the values of Kcr and Tcr
obtained and the corresponding proportional gains and integral gains obtained
for the current and the speed controllers. The effect of controller parameters
with stability analysis is given in Appendix 1.

Table 2.1 Kcr and Tcr values and the controller gains of the current and
speed controllers

Controller Kcr Tcr Proportional gain Integral gain


Current controller
208 0.0011s 93.6 1091
(q and d axes)
Speed controller 2.55 0.02862s 1.147 40.1

2.5.1.7 Sensor-less control

Using a sensor to encode the speed has several disadvantages like


increased cost, bigger size and lower reliability of the system. Figure 2.11
shows the block diagram of FOC of the PMSG without the sensor. In this
control, the speed is not measured with sensor but stator currents are used to
estimate the speed and position of the PMSG. The three phase currents of
stator are first transformed to d-q reference frame which are then transformed
to the reference frame. These currents are given to the position and the
speed estimation block, which estimate the speed and position of the rotor.
The estimated speed is compared with the reference speed and the error is
given to the PI speed controller which produces the reference q-axis current.
This current is compared with the actual q-axis current and the error is given
to q-axis current controller. The output of this controller is added with the
component II to get the q-axis voltage. The reference d-axis current (set at
zero) is compared with the actual d-axis current. The error is given to d-axis
current controller. Component I is subtracted from the output of the d-axis
48

current controller to produce the d-axis voltage. The voltages in d-q axes are
then transformed to axes which are given as reference voltages for the
SVM block. The speed and positions are estimated based on the flux linkage
estimation method. The various steps involved in flux linkage estimation
method are explained as follows.

2.5.1.8 Flux linkage estimation method

This method is used for estimating the position of the rotor. The
estimation of the rotor position is based on the estimation of the flux linkages
with the initial rotor position obtained from the flux linkages using the
inductances and the stator currents (Mora 2009). The rotor position estimation
is implemented in five steps.

Step 1: Estimation of the Stator Flux linkages

The phase currents and voltages are assumed to be measured at a


fixed sample time T. Equations (2.35) and (2.36) give the stator flux linkages
for the present instant

k T [v k 1 Rsi k ] k 1 (2.35)

k T [v k 1 Rsi k ] k 1 (2.36)

where, k is the present sampling interval and k-1 is the previous sampling
interval.

The currents i and i (i.e., the stator currents in the stationary reference
frame) are obtained by measuring the stator currents in phases a, b, and c
respectively.
49

IDC
GSC VDC
PMSG (VSC)
iabc
abc
dq
dq
i PWM
Position
and Speed
ids iqs er
Estimation
dq v SVM
PI
m
i*ds=0 &
PI PI PWM
*
* i qs
m
Component I
for max.
Wind speed (v w) Component II
power

Figure 2.11 Field oriented control of the PMSG without the sensor

Step 2: Estimation of the stator current

The stator currents are estimated using the flux linkages estimated
in the Step 1 and the predicted rotor position in Step 5. Equations (2.37) and
(2.38) can be solved to estimate the stator currents (i , i )

L sin 2 i L L cos 2 i m cos p (2.37)


p p

L sin 2 i L L cos 2 i m sin p (2.38)


p p
50

Ld Lq Lq Ld
where, L and L =0 for the surface mounted PMSG.
2 2

Step 3: Correction for the Position

In this step, the rotor position predicted in Step 5 is corrected


using the errors in stator currents ( i , i ), which are the differences
between the measured currents and the estimated currents in stationary
reference frames. Equations (2.39) and (2.40) give the current errors.

iq i cos i sin (2.39)


p p

i i sin i cos (2.40)


d p p

The corrected rotor position is obtained by adding the position error


to the predicted position, as given in Equation (2.41),

er (k ) p (k ) (k ) (2.41)

Lq i q
where, is the position error and , er and p are the corrected
m

and the predicted rotor positions, respectively. The determination of is


given in Appendix 1 (Mora 2009). Figure 2.12 shows the various steps
involved in the rotor position estimation.
51

Stator Currents, voltages,


flux linkages in Estimation of Stator Flux Linkage
stationary reference
frames

Current Estimation

Position Correction

Updating Flux Linkage

Rotor Position Position Prediction

Figure 2.12 Steps for the estimation of rotor position

Step 4: Updating the flux linkage

In this step, the fluxes are recalculated using the corrected rotor
position and the measured stator currents. Equations (2.42) and (2.43) express
the fluxes in discrete time for the surface mounted PMSG.

k Li k m cos er k (2.42)

k Li k m sin er k (2.43)

Step 5: Prediction of the rotor position

Considering the position to be a second-order polynomial, the


predicted position can be expressed by Equation (2.44),
52

2
p et ft g (2.44)

where, e, f, g are constants that are obtained by using a polynomial curve


fitting. The constants are determined by using three estimated positions at
sampling instants (k-2), (k-1) and k. The determination of the constants is
given in Appendix 2. (Cimpoeru 2010).

2.5.2 Grid Side Converter Control

To operate a wind energy system efficiently, the generator-side


controller alone is not enough. Though the generator-side converter controller
maintains the speed of the generator at an optimum speed to capture the
maximum power, the voltage of the DC-link capacitor varies with the
variation of the wind speed. This fluctuation directly affects the electrical grid
system, which leads to problems in weak grids. To prevent this problem, a
grid-side converter controller is needed. The grid side converter is also a VSC
which acts as an inverter that converts the DC-link voltage to an AC voltage
of desired frequency. The topology of the grid side converter is the same as
the generator side converter given in Figure 2.4. The input to the grid side
converter is the DC-link voltage. The grid side converter controller controls
the DC-link voltage of the system using a hysteresis current controller (HCC).
Figure 2.13 shows the block diagram of the grid side converter control with
the hysteresis current controller.
53

Grid Side
VDC Converter Grid
(VSC)

Pulses
iabcg
Hysteresis
Controller

i*abcg

V*DC 3- Reference
PI Current
Generator

Figure 2.13 Block diagram of the grid side converter control with the
hysteresis current controller

The DC-link voltage is compared with the reference DC voltage


and the error is given to the PI controller. The output of the PI controller is
fed to a three-phase reference current generator block, which generates the
three reference currents (i*abcg). These currents are compared with the grid
currents (iabcg) in the hysteresis controller to generate the PWM signals. The
converter connects the positive side of the DC-link to the grid when the
current produced in the hysteresis controller is lower than the reference
current, thereby increasing the grid currents. The converter connects the
negative side of the DC-link to the grid when the current produced in the
hysteresis controller is higher than the reference current, thereby decreasing
the grid currents.
54

2.5.2.1 Hysteresis current control (HCC)

The HCC maintains the DC-link voltage at the constant magnitude


by adjusting the power flow to the grid. The grid-side converter controller
uses three hysteresis current controllers (HCC) to generate the switching
pulses for the grid side converter. This technique is based on the on-line
PWM control that fixes the instantaneous inverter output voltage. The
function of the PWM current controller is to manipulate the actual output
currents to follow the reference currents. HCC is an instantaneous feedback
control that maintains the currents in a specified hysteresis band. The
differences between the reference currents and the actual currents determine
the switching signals for the switches of the GSC. The DC-link voltage is
adjusted by controlling the amount of current taken from the electrical grid.
The HCC controls the grid current by keeping it within the range of the
hysteresis band. It generates the PWM signal by comparing the grid current
with the lower and the higher limits of a band. It generates a high (1) when
the grid current reaches the lower limit of the band and generates a low (0)
when the grid current reaches the higher limit, generating the PWM signals.
The duty cycle of the PWM signals is adjusted by comparing the
instantaneous values of the grid currents with their reference currents. The
HCC technique has many advantages like being simple and robust, having
good dynamic response, having zero tracking errors and being unaffected by
load parameter changes (Baktiono 2012). Figure 2.14 shows the production of
gate signal with the hysteresis controller. The PI controller parameters for the
grid side controller are also designed using Ziegler Nichol’s second method of
tuning. The critical proportional gain (Kcr) obtained is 44.45 and the time
period of the sustained oscillation (Tcr) is 0.01s. The proportional gain for the
grid side PI controller is found to be 20 and the integral gain is found to be
2000.
55

ii (current) Hysteresis
Band

Actual current

Reference current

t (time)
t

Gate 1
Signal
0

Figure 2.14 Gate signal with the HCC

2.6 SIMULATION AND RESULTS

The sensor-less FOC controls the speed of the PMSG at the


reference speed or the set speed. Simulation is carried with the set speed
obtained from an optimum tip speed ratio, at which power coefficient has a
maximum value. When the PMSG and hence the turbine runs at this speed,
maximum power is captured from the wind.

The system with the back-back converters is simulated in


MATLAB for varying wind speeds. The simulation is carried out with the
Simulink models; the models used for simulation are given in Appendix 3.
The reference speed is obtained by using optimum tip speed ratio ( opt) and
the turbine rotor radius (R). The reference speed ( m*) equal to the the turbine
56

rotor speed ( w), turbine power (PT), turbine torque (TT) for three different
wind speeds (vw) (8.6 m/s, 12 m/s and 10 m/s) are calculated using
Equations (2.45) - (2.48) (Nantao Huang 2013).

* opt vw
Reference speed m w (2.45)
R

* 60
Reference speed in rpm N m
m (2.46)
2

1 3
Turbine power PT C p R 2 vw (2.47)
2

PT
Turbine torque TT (2.48)
w

1
1 x
c6
where, C p c1 c2 c3 c4 c5 e , c1=0.5, c2=116, c3=0.4,

1 1 0.035
c4=0, c5=5, c6=21, is pitch angle and . The torque
opt 0.08 1 3
should be negative for the machine to act as a generator. Since the machine is
a generator, power is delivered to the grid hence power is also negative. The
simulation is carried out based on the above values. The wind speed is
changed from 8.6 m/s to 12 m/s at 1 s and from 12 m/s to 10 m/s at 2 s. The
waveforms for the speed, torque, power, voltages of the PMSG, rotor
position, the DC-link voltage, the real and reactive powers fed to the grid are
shown for the changes in wind speed. The various parameters used for the
simulation are listed in Appendix 3.
57

Figures 2.15(a) shows the wind speed and Figure 2.15(b) shows the
reference and the measured speed of PMSG. Figures 2.16 and 2.17
respectively show, in the expanded scale, the reference speed and the
measured speed due to the increase and decrease in wind speeds.

Figure 2.15 (a) Wind speed

Reference speed

Measured speed

Figure 2.15 (b) The reference speed and the measured speed of the
PMSG for the increase in the wind speed at 1 s and the
decrease in the wind speed at 2 s
58

Reference speed

Measured speed

Figure 2.16 The reference speed and the measured speed of the PMSG,
in the expanded scale, for the increase in the wind speed at 1 s

Reference speed

Measured speed

Figure 2.17 The reference speed and the measured speed of the PMSG,
in the expanded scale, for the decrease in wind speed at 2 s

Figure 2.18 shows the electromagnetic torque of PMSG and the


turbine torque for the increase and decrease of the wind speed.
59

Turbine/ Mechanical Torque

Electromagnetic Torque

Figure 2.18 Electromagnetic torque of the PMSG and the turbine torque
for the increase in wind speed at 1 s and the decrease in
wind speed at 2 s

Figure 2.15(b) shows that the reference speed (found by Equation


2.46) is initially 537.4 rpm when the wind speed is 8.6 m/s and the PMSG
rotates at this speed. At 1 s, when the wind speed changes to 12 m/s, the
reference and the generator speeds change to 750 rpm. At 2 s, when the wind
speed changes to 10 m/s, the speeds change to 625 rpm. Figure 2.18 shows
that the turbine and the electromagnetic torques are initially -8 Nm at 1 s
which change to -13.38 Nm with increase in wind speed and the torques again
change to -10.72 Nm with decrease in wind speed at 2 s Hence, the generator
side controller is able to control the speed and torque of the PMSG.

Figure 2.19 shows the stator voltages of the PMSG and Figures
2.20, and 2.21 show the stator voltages, in the expanded scale, for the increase
and decrease of the wind speeds, respectively. Figure 2.20 shows that with
the increase in speed at 1 s, the stator voltages also increase while Figure 2.21
shows that with the decrease in speed at 2 s, the stator voltages also decrease.
Figure 2.22 shows the measured and estimated rotor positions. Figure 2.23
shows the generator power and the mechanical (turbine) power for the
increase and decrease in wind speeds.
Figure 2.19 Stator voltages of the PMSG for the increase in wind speed at 1 s and the decrease in wind speed at 2 s

60
61

Figure 2.20 Stator voltages of the PMSG, in the expanded scale, for the
increase in the wind speed at 1 s

Figure 2.21 Stator voltages of the PMSG, in the expanded scale, for the
decrease in the wind speed at 2 s

Estimated rotor position

Measured rotor position

Figure 2.22 Measured and estimated rotor positions


62

Turbine/Mechanical power

Generator power

Figure 2.23 Generator and turbine power for the increase in wind speed
at 1s and the decrease in wind speed at 2s

From Figure 2.23, the generator power is around -450W up to 1 s


for the rotor speed of 537.4 rpm. The power is changed to around -1060W at
1s when the rotor speed is 750 rpm. The power is changed to around -700 W
when the speed changes to 625 rpm. Figure 2.24 shows the reference and the
measured DC-link voltages, with the DC-link voltage set at 400 V. It is seen
that with the increase in the wind speed, the PMSG voltages also increases,
which results in the DC-link voltage increasing at 1 s and then settling to the
reference value at 1.05 s. When the wind speed decreases, the PMSG voltage
also decreases, thus the DC-link voltage decreases at 2 s and then settles to the
reference value at 2.05 s. It is evident from the Figure 2.24, that the DC-link
voltage is maintained constant at the set value of 400 V.

Figure 2.25 shows the real and reactive powers fed to the grid. The
reference real power is set at the generator power and the reference reactive
power is taken as zero. It is found from the Figure 2.25, that the real power fed
to the grid is almost equal to the generator power and the reactive power fed is
zero.
63

Reference DC voltage

Measured DC voltage

Figure 2.24 The Reference and the measured DC-link voltages for the
increase in the wind speed at 1 s and the decrease in the
wind speed at 2 s

Reactive power fed to grid (Q)

Real power fed to grid (P)

Figure 2.25 Real and the reactive powers fed into the grid

It is noted from simulated results that, the generator side converter


controller is able to control the speed and torque of PMSG and the grid side
converter controller is able to regulate the DC-link voltage and control the
grid powers.
64

2.7 SUMMARY

A wind energy conversion system with the PMSG and the


back-back converters is modeled. The PMSG, modeled in the synchronously
rotating reference frame, operates at variable speeds with the generator side
converter control. The sensor-less field oriented control technique is used for
the variable speed operation of the generator. This has the advantage of not
requiring an encoder for sensing speed, thus reducing the cost and
maintenance requirements. The modeled generator side converter with the
field oriented controller is able to control the speed and torque.

A grid side converter with a hysteresis current controller controls


the DC-link voltage. This voltage is controlled by controlling the current
drawn from the electrical grid. The hysteresis current controller controls the
grid current by keeping it within the range of the hysteresis band. The
designed grid side controller maintains the DC-link voltage at the set DC
voltage. The grid side controller also controls the real and reactive powers fed
to the grid.

The PI controllers for the generator side and the grid side converters
are designed using Zeigler-Nichol’s second method of tuning. It is found from
the simulated waveforms that the designed generator side controller is able to
control the speed and torque at the reference value while the designed grid
side controller is able to control the DC-link voltage, the real and reactive
powers fed to the grid at the set values.

At 1 s, when the wind speed changes from 8.6 m/s to 12 m/s,


reference and the generator speeds change from 537.4 rpm to 750 rpm, the
generator torque changes from -8 Nm to -13.38 Nm. At 2 s, when the wind
speed decreases from 12 m/s to 10 m/s , the reference and the generator speed
change from 750 rpm to 625 rpm, and the generator torque changes to
65

-10.72 Nm. Hence, the generator side controllers are able to control the speed
and torque of the PMSG. Similarly the grid side controller is able to control
the DC-link voltage. The DC-link voltage increases at 1s with the increase in
PMSG voltage (due to the increase in wind speed) and is brought back to the
reference of 400V at 1.05s. At 2 s, the DC-link voltage decreases due to the
decrease in the PMSG voltage and is brought back to the 400 V at 2.05 s. The
generator power and the real power fed to the grid are almost equal and the
reactive power fed to the grid is zero.

The next chapter deals with the various multilevel inverter


topologies. The Cascaded H-bridge (CHB) inverter (seven-level) with a single
DC source is considered and the various multicarrier PWM techniques are
discussed. A new selective harmonic elimination PWM (SHE-PWM)
technique is proposed. The new SHE-PWM has a lower THD when compared
to that of the conventional SHE-PWM.

You might also like