25.12.2012 Views

Damping of Wind Turbine Tower Oscillations through Rotor Speed ...

Damping of Wind Turbine Tower Oscillations through Rotor Speed ...

Damping of Wind Turbine Tower Oscillations through Rotor Speed ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

March 29 – April 1<br />

International Conference on<br />

Ecologic Vehicles & Renewable Energies<br />

<strong>Damping</strong> <strong>of</strong> <strong>Wind</strong> <strong>Turbine</strong> <strong>Tower</strong> <strong>Oscillations</strong><br />

<strong>through</strong> <strong>Rotor</strong> <strong>Speed</strong> Control<br />

Copyright © 2007 MC2D & MITI<br />

Mate Jelavić, Nedjeljko Perić, Ivan Petrović<br />

Faculty <strong>of</strong> Electrical Engineering and Computing<br />

University <strong>of</strong> Zagreb, 10000 Zagreb, Croatia<br />

E-mail: mate.jelavic@fer.hr<br />

nedjeljko.peric@fer.hr<br />

ivan.petrovic@fer.hr<br />

Abstract: To enable wind turbines to produce power under great variety <strong>of</strong> wind conditions a<br />

sophisticated control system is needed. <strong>Wind</strong> turbine system is highly nonlinear and its dynamic<br />

changes rapidly with the change <strong>of</strong> wind speed. Besides this, wind turbine mechanical structure is very<br />

flexible due to its great height and tends to oscillate. All this makes the design <strong>of</strong> wind turbine control<br />

system very demanding task. In this paper three controller design methods are compared. It is shown<br />

how with adequate controller design it becomes possible to control the wind turbine under various<br />

operating conditions reducing the structural oscillations at the same time.<br />

Keywords: <strong>Wind</strong> turbine, pitch control, tower oscillations, pole placement, full state feedback<br />

1. Introduction<br />

Modern wind turbines have to operate in<br />

wide range <strong>of</strong> operating conditions determined<br />

primarily by wind speed. To make it possible<br />

for wind turbine to produce power in such a<br />

variety <strong>of</strong> operating conditions a sophisticated<br />

control system is needed that will account for<br />

changes in operating conditions and<br />

accompanying changes in wind turbine<br />

dynamics.<br />

<strong>Wind</strong> power or the power <strong>of</strong> air that moves<br />

at speed vw over the area swept by turbine rotor<br />

with radius R is given by [1]:<br />

1 2 3<br />

Pw = ρairR πvw,<br />

(1)<br />

2<br />

where ρ air is density <strong>of</strong> air. From expression<br />

(1) it is clear that wind energy increases rapidly<br />

with increase in wind speed. This results in two<br />

very different operation regions <strong>of</strong> wind<br />

turbine, each <strong>of</strong> them placing specific demands<br />

upon control system. During weak winds<br />

power contained in the wind is lower than the<br />

rated power output <strong>of</strong> wind turbine generator.<br />

Therefore, the main task <strong>of</strong> the control system<br />

in this region is to maximize wind turbine<br />

power output by maximizing wind energy<br />

capture. It can be shown [1] that for each value<br />

<strong>of</strong> wind speed energy conversion efficiency is<br />

maximal for only one particular value <strong>of</strong> rotor<br />

speed. Since modern wind turbines are<br />

connected to grid using AC-DC-AC frequency<br />

converters, generator frequency is decoupled<br />

from grid frequency what enables variable<br />

speed operation. Therefore it becomes possible<br />

to vary the rotor speed and to maintain optimal<br />

energy conversion during varying wind speeds.<br />

On the other hand, during strong winds power<br />

<strong>of</strong> the wind is greater than the rated power<br />

output <strong>of</strong> wind turbine generator. Therefore,<br />

the wind energy conversion has to be<br />

constrained in this region to assure generator<br />

operation without overloading. Very efficient<br />

method for constraining wind energy<br />

conversion is pitching the rotor blades around<br />

their longitudinal axis what deteriorates their<br />

aerodynamic efficiency and therefore only a<br />

part <strong>of</strong> wind energy is used for driving the<br />

generator.


The borderline between two operation<br />

regions is the lowest wind speed at which<br />

turbine generator reaches its rated power<br />

output. This wind speed is termed rated wind<br />

speed [1]. Hence the wind turbine operation<br />

regions are known as below and above rated<br />

regions. In this paper we focus on the above<br />

rated operation region and describe design <strong>of</strong><br />

controllers used in that region.<br />

The paper is organized as follows. Section 2<br />

gives a general description <strong>of</strong> the wind turbine<br />

control system. Section 3 introduces<br />

mathematical model <strong>of</strong> the wind turbine that is<br />

used for controller design. Section 4 describes<br />

the design <strong>of</strong> PID controller which is nowadays<br />

standard solution for wind turbine control. In<br />

section 5 and 6 two alternative control concepts<br />

are proposed and their performances are<br />

compared with the performance <strong>of</strong> PID<br />

controller. Conclusions are given in section 7.<br />

2. <strong>Wind</strong> turbine control system<br />

The main task <strong>of</strong> wind turbine control system<br />

is to obtain continuous power production under<br />

operating conditions determined by various<br />

wind speeds. As turbine power is directly<br />

proportional to its speed, power control can be<br />

done by controlling turbine speed. The<br />

principle scheme <strong>of</strong> wind turbine speed control<br />

system is shown in fig. 1. As it can be seen in<br />

this figure turbine speed can be influenced and<br />

Figure 1: Principle scheme <strong>of</strong> wind turbine control<br />

system.<br />

thus controlled by two means – by generator<br />

electromagnetic torque Mg which opposes rotor<br />

driving torque Mr and by pitch angle β which<br />

alters the wind energy conversion. For this<br />

reason turbine speed control system consists <strong>of</strong><br />

two control loops: torque control loop and pitch<br />

control loop. Those control loops operate<br />

simultaneously but depending on operation<br />

region one <strong>of</strong> them is dominant. In the below<br />

rated operation region the torque control loop is<br />

used to control turbine speed to values that will<br />

result in maximal wind power capture. This<br />

control loop is not in the scope <strong>of</strong> this paper.<br />

Details on its specifics can be found in e.g. [2].<br />

In the above rated region this control loop just<br />

holds generator torque at its rated value.<br />

The pitch control loop is used for setting the<br />

adequate pitch angle that will keep turbine<br />

speed at its reference value under all operating<br />

conditions determined by various winds. Below<br />

rated wind speed this loop sets pitch angle to<br />

value that assures maximal wind power capture<br />

which is usually around 0 o . In this paper we<br />

assume that all blades have the same pitch<br />

angle what is known as "collective pitch".<br />

Controller in this loop, although used to control<br />

turbine speed, is commonly termed pitch<br />

controller. Blade positioning is mostly done<br />

using electrical servo drives that rotate blades<br />

by means <strong>of</strong> gearboxes and slewing rings.<br />

Position control <strong>of</strong> servo drives is usually<br />

achieved using frequency converters. This<br />

control loop design is rather simple and is not<br />

in the scope <strong>of</strong> the paper. For simulation<br />

analysis it is modeled as second order system:<br />

SD<br />

( )<br />

G s<br />

( s)<br />

( )<br />

β<br />

ω<br />

= =<br />

β ζω ω<br />

ref s<br />

2<br />

s + 2 β<br />

2<br />

nβ<br />

nβs+ 2<br />

nβ<br />

. (2)<br />

In this paper we use ω nβ =6.28 rad/s and<br />

ζ β = 2/2 what is fairly good approximation<br />

<strong>of</strong> real pitch positioning system.<br />

3. <strong>Wind</strong> turbine model<br />

The first step in control system design is<br />

construction <strong>of</strong> a suitable process model for<br />

wind turbine in scope. <strong>Wind</strong> energy<br />

conversion process is highly nonlinear and<br />

difficult for mathematical description. It can be<br />

described quite well using combined blade<br />

element and momentum theory [1]. However<br />

this approach yields implicit mathematical<br />

expressions that can only be solved iteratively.<br />

This form, although very common in<br />

simulation tools, is not suitable for controller<br />

design. So in this paper we use another<br />

somewhat simplified mathematical model that<br />

is usual in the literature dealing with controller<br />

design. Here we describe it briefly, while<br />

details on it can be found in [3] and [4].<br />

Power contained in wind given by<br />

expression (1) can never be completely<br />

transformed into wind turbine power and<br />

afterwards into electrical power. The amount <strong>of</strong><br />

wind power that is converted into turbine<br />

power Pr can be described by means <strong>of</strong><br />

performance coefficient CP [1]:


Pr = Pw⋅ CP.<br />

(3)<br />

The theoretical maximum for CP is<br />

determined by the Betz' law [1] and equals<br />

16/27. In practice wind turbines don't reach this<br />

limit but approach the value <strong>of</strong> 0.5 at best.<br />

CP is not a constant parameter but its value is<br />

dependant on wind speed v w , rotor speed ω<br />

and blade pitch angle β . <strong>Wind</strong> speed and rotor<br />

speed are usually bound together introducing<br />

parameter λ that is called tip speed ratio which<br />

represents the ratio between blade tip speed and<br />

wind speed [1]:<br />

ωR<br />

λ = . (4)<br />

v<br />

Typical dependence <strong>of</strong> performance<br />

coefficient upon tip speed ratio with pitch angle<br />

used as a parameter is shown in fig. 2.<br />

Figure 2: Performance coefficient as a function <strong>of</strong><br />

tip speed ratio.<br />

Aerodynamic torque Mr that drives wind<br />

turbine rotor and thus generator is given by:<br />

M<br />

r<br />

w<br />

( , )<br />

2 3<br />

P 1 ρairR πv r<br />

wCPλ β<br />

= = . (5)<br />

ω 2 ω<br />

Rearrangement <strong>of</strong> expression (5) yields:<br />

M<br />

r<br />

( , )<br />

3 2<br />

1 ρairR πvwCPλ β<br />

= . (6)<br />

2 λ<br />

A quotient <strong>of</strong> performance coefficient and tip<br />

speed ratio forms a new dimensionless<br />

parameter that is known as torque coefficient<br />

CQ [1]:<br />

C<br />

Q<br />

( λβ , )<br />

C<br />

( λ, β )<br />

P<br />

= . (7)<br />

Having aerodynamic torque calculated<br />

according to (6) rotor speed can easily be found<br />

using principle equation <strong>of</strong> motion:<br />

λ<br />

J � ω = M −M −M<br />

, (8)<br />

t r g l<br />

where Mg is generator electromagnetic torque,<br />

Jt is total moment <strong>of</strong> inertia <strong>of</strong> rotor and<br />

generator while Ml is loss torque. Loss torque<br />

Ml, caused mostly by friction is rather small<br />

and will be neglected here.<br />

In this paper we consider wind turbine with<br />

generator that is directly coupled with turbine<br />

rotor. This turbine setting known as direct drive<br />

system uses synchronous multipole generator<br />

that rotates at small speed <strong>of</strong> turbine rotor.<br />

Since rotor and generator speeds are the same<br />

no distinction between them is made<br />

<strong>through</strong>out the paper. Because there is no<br />

gearbox between rotor and generator their<br />

moments <strong>of</strong> inertia can just be summed<br />

together in order to calculate total moment <strong>of</strong><br />

inertia Jt. The coupling <strong>of</strong> rotor to the generator<br />

in direct drive solutions is very stiff and it can<br />

be considered as rigid thus removing any<br />

torsional oscillations what simplifies the<br />

control system design.<br />

Before going further an important issue has<br />

to be addressed. Namely, expressions (4), (5)<br />

and (6) in this form would be valid only for<br />

structure with rigid tower and blades. In real<br />

situation the absolute wind speed v w in<br />

mentioned expressions has to be replaced by<br />

wind speed that is "seen" by rotor blades. This<br />

wind speed seen by the rotor is the resultant <strong>of</strong><br />

three factors: absolute wind speed v w , speed<br />

<strong>of</strong> the tower movement perpendicular to wind<br />

speed (i.e. tower nodding speed) x� t and speed<br />

<strong>of</strong> blade movement perpendicular to wind<br />

speed (i.e. speed <strong>of</strong> blade flapwise movement).<br />

Influence <strong>of</strong> tower nodding on wind turbine<br />

control is much more pronounced than<br />

influence <strong>of</strong> blade flapwise movement.<br />

Therefore we focus only on tower nodding<br />

considering rotor blades as rigid.<br />

<strong>Tower</strong> nodding originates from the fact that<br />

wind turbine tower is very lightly damped<br />

structure due to its great height (more than 100<br />

meters in modern wind turbines) and need for<br />

moderate mass. To model the wind turbine<br />

tower precisely we would have to use model<br />

with distributed parameters and to describe it<br />

in terms <strong>of</strong> mass and stiffness distribution.<br />

Such a model wouldn't be very suitable for<br />

controller design so it has to be substituted by


model with concentrated parameters. This can<br />

be done using modal analysis that is very<br />

common tool in wind turbine analysis [1], [3].<br />

It describes a complex oscillatory structure as<br />

a composition <strong>of</strong> several simple oscillatory<br />

systems each <strong>of</strong> them being described by<br />

means <strong>of</strong> mass, stiffness and damping. By this<br />

representation complex tower oscillations are<br />

seen as a sum <strong>of</strong> many simple oscillations<br />

characterized by their modal frequencies<br />

which are one <strong>of</strong> the most important structural<br />

properties <strong>of</strong> wind turbine. It has been shown<br />

in practice [5] that fairly good modeling <strong>of</strong><br />

wind turbine tower nodding can be achieved<br />

using two modal frequencies (two modes).<br />

Since we are here primarily interested in<br />

building model suitable for controller design<br />

we use only the first modal frequency. The<br />

justification for this lies in the fact that for the<br />

turbine in scope second modal frequency is<br />

more than 6 times greater than the first modal<br />

frequency and therefore falls out <strong>of</strong> the<br />

controller frequency bandwidth.<br />

By using only one modal frequency tower<br />

dynamics can be described as:<br />

Mx �� + Dx� + Cx = F , (9)<br />

t t t<br />

where M, D, and C are modal mass damping<br />

and stiffness respectively and F is the<br />

generalized force that is originated by wind<br />

and that causes wind turbine tower<br />

oscillations. <strong>Tower</strong> modal properties in<br />

expression (9) are related to first tower modal<br />

frequency ω0t as follows [4]:<br />

D= 2 ζ ω ⋅M,<br />

t 0t<br />

2<br />

0t<br />

( ω )<br />

C = ⋅M,<br />

(10)<br />

where ζ t is structural damping. For steel<br />

structure structural damping is mostly set to<br />

0.005 [4]. Modal mass M can be calculated as<br />

[1]:<br />

ht<br />

0<br />

( ) ( ) 2<br />

φ<br />

M = ∫ m h h dh,<br />

(11)<br />

where ht is the height <strong>of</strong> the tower, m(h) is the<br />

mass distribution along the tower and φ ( h)<br />

is<br />

the tower's first mode shape. Note that actual<br />

distribution <strong>of</strong> mass along tower has to be<br />

modified in order to include mass <strong>of</strong> the rotor<br />

and the nacelle which is assumed to be<br />

concentrated at the tower top.<br />

Driving force F is mostly the rotor thrust force<br />

Ft caused by wind. It can be shown [4] that<br />

thrust force, similar to aerodynamic torque,<br />

depends upon wind speed, rotor speed and<br />

pitch angle. So similarly to (6) it can be<br />

expressed as [4]:<br />

1 2 2<br />

Ft = ρairR πvwCt( λ, β)<br />

, (12)<br />

2<br />

where Ct is so called thrust coefficient.<br />

Expressions (6), (8), (9) and (12) form the<br />

simplified nonlinear model <strong>of</strong> wind turbine<br />

that is used in the following sections for<br />

controller design. Model is summarized below<br />

taking into account the fact that wind speed<br />

seen by the rotor is a sum <strong>of</strong> wind speed and<br />

tower nodding speed:<br />

J � tω= Mr −Mg −Ml,<br />

1 3<br />

2<br />

Mr = ρairR πCQ( λ, β)<br />

⋅( vw −x�t)<br />

,<br />

2<br />

. (13)<br />

1 2<br />

2<br />

Ft = ρairR πCt( λ, β)<br />

⋅( vw −x�t)<br />

,<br />

2<br />

F = Mx �� + Dx�+ Cx.<br />

t t t t<br />

Torque and thrust coefficients Cq and Ct are<br />

usually provided by wind turbine blade<br />

manufacturers or can be calculated using<br />

pr<strong>of</strong>essional simulation tools.<br />

4. PID Controller<br />

The PID controllers are still by far the most<br />

used controllers for wind turbine speed and<br />

power control. This is due to their simplicity<br />

and rather high robustness. The small number <strong>of</strong><br />

parameters makes possible for designer to<br />

quickly arrive at satisfactory, although in many<br />

cases suboptimal, system behavior. As stated<br />

before wind turbine dynamics change in<br />

nonlinear fashion with change in wind speed.<br />

To control the wind turbine with linear<br />

controllers gain scheduling has to be used [3].<br />

Although it seems straightforward to use wind<br />

speed as scheduling criterion this is not an<br />

appropriate solution since the wind speed is not<br />

measured fast and accurately enough. Therefore<br />

measured pitch angle is usually used as<br />

scheduling variable.<br />

For the design <strong>of</strong> a PID controller using<br />

analytical methods process model (13) has to<br />

be linearised around chosen operating point.<br />

After linearization <strong>of</strong> expressions (13) and


transition to Laplace domain simple algebraic<br />

manipulations yield transfer functions that are<br />

needed for controller design. Those transfer<br />

functions are:<br />

and<br />

( )<br />

G s<br />

β<br />

w<br />

( )<br />

G s<br />

Δω<br />

=<br />

Δβ<br />

w<br />

( s)<br />

( s)<br />

( s)<br />

( )<br />

Δω<br />

=<br />

Δv<br />

s<br />

(14)<br />

. (15)<br />

These transfer functions are <strong>of</strong> the third order.<br />

A good insight in system properties <strong>of</strong> the<br />

wind turbine can be gained if we examine<br />

frequency characteristics <strong>of</strong> transfer function<br />

(14) shown in fig. 3.<br />

Figure 3: Frequency characteristics <strong>of</strong> G β .<br />

It can be observed that frequency<br />

characteristics <strong>of</strong> transfer function (14) at<br />

tower modal frequency exhibits magnitude and<br />

phase drop. Similar phenomena are present in<br />

frequency characteristics <strong>of</strong> (15) as well. This<br />

fact makes the pitch controller design very<br />

difficult. Physical explanation for observed<br />

effects becomes clear from the following<br />

analysis. Change in wind speed causes change<br />

in rotor speed what requires controller action<br />

and pitching <strong>of</strong> the blades in order to regulate<br />

the rotor speed to its rated value. Pitching the<br />

rotor blades, besides the aerodynamic torque,<br />

alters the thrust force significantly. Thrust<br />

force, according to (13), causes change in wind<br />

turbine tower top speed and thus the wind<br />

speed seen by the rotor is changed. This alters<br />

the aerodynamic conversion and in this way a<br />

feedback is formed. For this reason wind<br />

turbine can easily be driven into oscillatory<br />

behavior if the pitch controller is not designed<br />

properly.<br />

To prevent the pitch controller from driving the<br />

wind turbine into oscillatory behavior it must be<br />

assured that system frequency bandwidth is<br />

below the first tower modal frequency.<br />

Moreover, sufficiently small magnitude is<br />

required at the first modal frequency. As it can<br />

be seen from fig. 3. the first modal frequency <strong>of</strong><br />

the turbine in scope is 3 rad/s so a bandwidth <strong>of</strong><br />

1 rad/s was chosen. PID controller was designed<br />

to assure phase margin <strong>of</strong> around 60 o what gives<br />

satisfactory behavior <strong>of</strong> the system.<br />

To fully explore the system behavior with<br />

chosen controller simulation tool GH Bladed<br />

was used. GH Bladed is pr<strong>of</strong>essional simulation<br />

package designed for wind turbine simulations<br />

and load calculations [5]. It relies upon very<br />

complex mathematical model based on<br />

combined blade element and momentum theory<br />

[1]. Structural properties <strong>of</strong> the wind turbine are<br />

modeled in detail and inertial and gravitational<br />

loads are taken into account along with<br />

aerodynamic ones. Extensive testing showed<br />

that simulation results obtained in Bladed are in<br />

accordance with measurements taken on actual<br />

wind turbines what was recognized by major<br />

standardization and certification institutions<br />

(e.g. Germanischer Llyod).<br />

To model the structural properties <strong>of</strong> explored<br />

turbine many modes are used [5]. <strong>Tower</strong><br />

nodding is modeled with two modes as well as<br />

tower naying (tower side-side motion). <strong>Rotor</strong><br />

blades' motion in flapwise direction is modeled<br />

with 6 modal frequencies while blades' motion<br />

in edgewise direction (displacement <strong>of</strong> the rotor<br />

blades in the plane <strong>of</strong> rotation) is modeled with<br />

5 modal frequencies. <strong>Wind</strong> shear and tower<br />

shadow are included in the model as well. PID<br />

controller designed based on linearised model<br />

(14) and (15) was implemented in C and<br />

included as external discrete time controller. In<br />

that way we could use Bladed for controller<br />

testing.<br />

In the following figures behavior <strong>of</strong> the system<br />

when PID controller is used for rotor speed<br />

control is shown for one representative<br />

operating point ( v w = 15 m/s). <strong>Wind</strong> that was<br />

used for simulation observed positive and<br />

negative stepwise change shown in fig. 4.<br />

Responses <strong>of</strong> rotor speed, pitch angle and tower<br />

top displacement are shown in figs. 5. 6. and 7.<br />

respectively. From these figures it can be seen<br />

that rotor speed is well regulated and quickly<br />

compensated for the influences <strong>of</strong> wind speed<br />

changes. The pitch control actions are moderate<br />

without any oscillations. Similar results were<br />

obtained for all operating points <strong>through</strong>out<br />

wind turbine operating range.


<strong>Wind</strong> speed [m/s]<br />

18<br />

17<br />

16<br />

15<br />

14<br />

13<br />

12<br />

0 10 20 30 40 50 60 70<br />

t [s]<br />

Figure 4: <strong>Wind</strong> speed used for simulation in Bladed.<br />

<strong>Rotor</strong> speed [rpm]<br />

25<br />

24.5<br />

24<br />

23.5<br />

23<br />

22.5<br />

22<br />

0 10 20 30 40 50 60 70<br />

t [s]<br />

Figure 5: Response <strong>of</strong> rotor speed <strong>of</strong> the system<br />

controlled with PID controller.<br />

Pitch angle [deg]<br />

16<br />

15<br />

14<br />

13<br />

12<br />

11<br />

10<br />

9<br />

8<br />

7<br />

0 10 20 30 40 50 60 70<br />

t [s]<br />

Figure 6: Response <strong>of</strong> pitch angle <strong>of</strong> the system<br />

controlled with PID controller.<br />

<strong>Tower</strong> top displacement [m]<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0<br />

-0.02<br />

-0.04<br />

-0.06<br />

-0.08<br />

0 10 20 30 40 50 60 70<br />

t [s]<br />

Figure 7: Response <strong>of</strong> tower top displacement <strong>of</strong> the<br />

system controlled with PID controller.<br />

While rotor speed and pitch angle behavior is<br />

satisfactory, attention should be paid to the<br />

tower top oscillations shown in fig. 7. It can be<br />

observed that tower top experiences lightly<br />

damped oscillatory behavior. These lasting<br />

oscillations, although small in magnitude,<br />

contribute to material fatigue and can lead to<br />

structure premature failure. So in the following<br />

sections two methods for controller design that<br />

aim at reduction <strong>of</strong> tower top oscillations are<br />

proposed.<br />

5. Input-output pole placement<br />

controller<br />

In this section the pitch controller with inputoutput<br />

structure is designed using well known<br />

pole placement method [6]. As design objective<br />

in this method the desired behavior <strong>of</strong> the<br />

closed loop system has to be chosen. The pitch<br />

controller GPC is then designed to assure that<br />

closed loop system behaves in the chosen<br />

manner. Transfer functions (14) and (15) form<br />

the linearised wind turbine model that can be<br />

described with the principle scheme shown in<br />

fig. 8.<br />

Figure 8: Principle scheme <strong>of</strong> the linearised wind<br />

turbine model.<br />

From the principle scheme given in fig. 8 closed<br />

loop transfer function with respect to wind<br />

speed change can be derived:<br />

G<br />

CL_ w<br />

Δω()<br />

s<br />

Gw() s<br />

= =<br />

. (16)<br />

Δ v () s 1 + G () s G s G () s<br />

w CL PC SD<br />

( )<br />

The pitch controller GPC has to assure that<br />

closed loop transfer function (16) is equal to the<br />

chosen model transfer function Gm:<br />

( ) !<br />

G s = G () s . (17)<br />

CL _ w m<br />

β


From (16) and (17) it follows that pitch<br />

controller has a form <strong>of</strong> :<br />

1 Gw() s −Gm()<br />

s<br />

GPC() s = ⋅ . (18)<br />

G s G () s G () s<br />

( )<br />

SD β<br />

m<br />

It should be noted that the above expression<br />

<strong>of</strong>ten needs to be modified to assure that the<br />

controller is causal. Details on this design<br />

method can be found in [6].<br />

The crucial step in controller design using<br />

described method is the choice <strong>of</strong> model<br />

transfer function Gm. It is important to choose<br />

model transfer that is achievable for the system<br />

in scope and that assures its satisfactory<br />

behavior. One possibility that is very common<br />

in the literature is the use <strong>of</strong> standard forms<br />

such as Butterworth or binomial form.<br />

However, these forms can be rather difficult to<br />

relate to system physical properties. So another<br />

approach is used in this paper. As previously<br />

said our primary goal is the reduction <strong>of</strong> tower<br />

oscillations. <strong>Tower</strong> oscillations can be reduced<br />

if the tower modal damping increases what<br />

would require change in tower structural<br />

parameters such as mass and stiffness<br />

distributions. Our goal is to achieve similar<br />

increase <strong>of</strong> tower damping by means <strong>of</strong> pitch<br />

controller actions without change in tower<br />

structural parameters. In that sense as a<br />

controller design objective we set a desired<br />

increase <strong>of</strong> tower modal damping. In other<br />

words tower modal damping D in the last<br />

expression in (13) is replaced with desired<br />

modal damping D' thus forming system model<br />

with new set <strong>of</strong> parameters. This model is<br />

linearised and transfer functions (14) and (15)<br />

are calculated. Using calculated transfer<br />

functions PID controller is designed following<br />

the guidelines described in section 4. Having<br />

the controller designed it is possible to calculate<br />

closed loop transfer function (16). This closed<br />

loop transfer function, obtained using system<br />

with increased damping and PID controller is<br />

then regarded as desired model transfer<br />

functions Gm for the real system. In other words<br />

our goal is to design a controller that assures<br />

that real system behaves as its tower damping<br />

has increased. The system response with Pole<br />

placement controller was simulated in Bladed<br />

using previously described full featured model.<br />

<strong>Wind</strong> used for simulation was the same as in<br />

section 4. Simulation results are given in the<br />

figures 9-11. From these figures it can be seen<br />

that pole placement controller achieves almost<br />

the same regulation <strong>of</strong> rotor speed as PID<br />

<strong>Rotor</strong> speed [rpm]<br />

25<br />

24.5<br />

24<br />

23.5<br />

23<br />

22.5<br />

22<br />

0 10 20 30 40 50 60 70<br />

t [s]<br />

Figure 9: Response <strong>of</strong> rotor speed <strong>of</strong> the system<br />

controlled with input-output pole placement<br />

controller.<br />

Pitch angle [deg]<br />

16<br />

15<br />

14<br />

13<br />

12<br />

11<br />

10<br />

9<br />

8<br />

7<br />

0 10 20 30 40 50 60 70<br />

t [s]<br />

Figure 10: Response <strong>of</strong> pitch angle <strong>of</strong> the system<br />

controlled with input-output pole placement<br />

controller.<br />

<strong>Tower</strong> top displacement [m]<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0<br />

-0.02<br />

-0.04<br />

-0.06<br />

0 10 20 30 40 50 60 70<br />

t [s]<br />

Figure 11: Response <strong>of</strong> tower top displacement <strong>of</strong><br />

the system controlled with input-output pole<br />

placement controller.<br />

controller does. This was to be expected since<br />

the PID controller is the "core" <strong>of</strong> the pole<br />

placement controller design. At the same time<br />

tower oscillations are more damped. Since the<br />

tower structure remains the same this damping<br />

is achieved by higher pitch control activity than<br />

in the case <strong>of</strong> PID controller.<br />

The question that naturally appears is the limit<br />

to which extend the tower oscillations can be<br />

damped in this way. Answer to this question can


e obtained analyzing fig. 12. This figure shows<br />

comparison <strong>of</strong> Bode plots <strong>of</strong> controllers<br />

designed with different values <strong>of</strong> desired tower<br />

damping D'.<br />

Figure 12: Bode plots <strong>of</strong> input-output pole placement<br />

controllers designed to achieve different values <strong>of</strong><br />

tower damping D'.<br />

From fig. 12 it becomes clear that increased<br />

tower damping results in high pass controller<br />

behavior. This has a consequence <strong>of</strong> increased<br />

pitch activity which in turn results in additional<br />

oscillations and finally cancels out the<br />

advantages <strong>of</strong> proposed design methodology. So<br />

a trade<strong>of</strong>f between desired increase in tower<br />

damping and pitch activity has to be made.<br />

6. Full state feedback controller<br />

The input-output pole placement controller<br />

described in section 5 has shown some<br />

promising results but its ability to damp the<br />

tower oscillations is limited. Key cause for this<br />

is the fact that it doesn't use information about<br />

actual tower oscillations but only rotor speed<br />

feedback. <strong>Damping</strong> <strong>of</strong> tower oscillations by<br />

addition <strong>of</strong> tower top speed feedback to PID<br />

controller is proposed in [1]. <strong>Tower</strong> top speed<br />

measurement can be obtained from tower top<br />

acceleration measured by accelerometers that<br />

are nowadays almost a standard part <strong>of</strong> wind<br />

turbine control system. In our approach we use<br />

slightly different methodology. Instead <strong>of</strong><br />

extended PID controller we use full state<br />

feedback controller designed using pole<br />

placement method. Desired closed loop<br />

behavior is chosen in the same way as for<br />

described pole placement controller. For this<br />

purpose process model (13) has to be rewritten<br />

in the state space form:<br />

x� = A⋅ x+ B⋅u, y = C⋅ x+ D⋅u. (19)<br />

State variables used for system description and<br />

control are rotor speed ω , rotor acceleration<br />

ω� , tower top speed x� t and tower top<br />

acceleration x�� t , while system inputs are wind<br />

speed v w , pitch angle β and generator torque<br />

M g :<br />

⎡ω⎤ ⎢ ⎡ v ⎤ w<br />

� ω<br />

⎥<br />

⎢ ⎥<br />

x = ⎢ ⎥,<br />

u = β<br />

x ⎢ ⎥.<br />

(20)<br />

⎢�⎥ t<br />

⎢ ⎥ ⎢M⎥ g<br />

x ⎣ ⎦<br />

⎣��t ⎦<br />

<strong>Rotor</strong> speed and tower acceleration are<br />

measured variables while other two states are<br />

derived from them. Using Ackermann's formula<br />

[6] vector <strong>of</strong> feedback gains for selected states<br />

can be calculated. System with such a controller<br />

was tested in Bladed using the same wind<br />

stepwise change as in sections 4 and 5.<br />

Simulation results are shown in the figures 13-<br />

15. From these figures it can be seen that full<br />

state feedback controller maintains good rotor<br />

speed regulation while at the same time<br />

achieving better damping <strong>of</strong> the tower<br />

oscillations and practically removing the<br />

oscillatory movement <strong>of</strong> the tower. The pitch<br />

activity in this case increases considerably so a<br />

trade<strong>of</strong>f between tower oscillations' damping and<br />

pitch activity is necessary.<br />

To examine the behavior <strong>of</strong> three described<br />

controllers in more realistic conditions the<br />

system behavior in 3D turbulent wind field was<br />

simulated in Bladed. Further information about<br />

simulations <strong>of</strong> 3D turbulent wind fields can be<br />

found in e.g. [7]. Let's just mention here that<br />

turbulent wind field was generated in Bladed<br />

<strong>Rotor</strong> speed [rpm]<br />

25<br />

24.5<br />

24<br />

23.5<br />

23<br />

22.5<br />

22<br />

0 10 20 30 40 50 60 70<br />

t [s]<br />

Figure 13: Response <strong>of</strong> rotor speed <strong>of</strong> the system<br />

controlled with full state feedback controller.


Pitch angle [deg]<br />

16<br />

15<br />

14<br />

13<br />

12<br />

11<br />

10<br />

9<br />

8<br />

7<br />

0 10 20 30 40 50 60 70<br />

t [s]<br />

Figure 14: Response <strong>of</strong> pitch angle <strong>of</strong> the system<br />

controlled with full state feedback controller.<br />

<strong>Tower</strong> top displacement [m]<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0<br />

-0.02<br />

-0.04<br />

0 10 20 30 40 50 60 70<br />

t [s]<br />

Figure 15: Response <strong>of</strong> tower top displacement <strong>of</strong><br />

the system controlled with full state feedback<br />

controller.<br />

using Kaimal spectrum recommended by the<br />

international standards for wind turbine design<br />

[8]. To gain valid information about system<br />

behavior in 3D turbulent conditions it is<br />

necessary for the simulations to last for at least<br />

10 minutes what generates lots <strong>of</strong> data.<br />

Therefore, simulation results are not presented<br />

here due to limited space. Simulations in 3D<br />

turbulent wind field have shown that inputoutput<br />

pole placement controller in turbulent<br />

conditions achieves only modest improvement <strong>of</strong><br />

tower oscillations when compared to the PID<br />

controller. The reason for this is the mentioned<br />

fact that input-output controller doesn't use<br />

information about actual tower oscillations. On<br />

the other hand full state feedback controller<br />

maintains shown ability to reduce tower<br />

oscillations even under turbulent winds. The cost<br />

<strong>of</strong> that is increased pitch activity what must be<br />

taken into account to avoid pitch system<br />

excessive wear. Simple design method proposed<br />

in this paper makes it possible to easily change<br />

the desired increase in tower damping and to<br />

achieve a trade<strong>of</strong>f between reduction <strong>of</strong> tower<br />

oscillations and increase <strong>of</strong> pitch activity.<br />

7. Conclusion<br />

The control system for variable speed pitch<br />

controlled wind turbine is presented. The wind<br />

turbine system is highly nonlinear and its<br />

parameters change significantly with change <strong>of</strong><br />

wind speed. Furthermore wind turbine<br />

mechanical structure is very flexible and can<br />

easily be driven into oscillatory behavior. All<br />

this makes the controller design a very<br />

demanding task. In this paper three methods<br />

for wind turbine pitch controller design are<br />

compared. It is shown that classic PID<br />

controller can assure good rotor speed<br />

regulation but tower oscillations are very<br />

pronounced. To reduce these undesired<br />

oscillations two alternative control structures<br />

are investigated: input-output pole placement<br />

controller and full state feedback controller.<br />

The design objective for these controllers,<br />

besides good rotor speed regulation, was the<br />

increase <strong>of</strong> tower damping. Both controllers<br />

have shown that owing to increased pitch<br />

control activity it becomes possible to damp<br />

the tower oscillations. To test the controllers'<br />

performances in more realistic conditions<br />

simulation under 3D turbulent wind field were<br />

conducted in Bladed. It was observed that<br />

under such conditions only slight reduction <strong>of</strong><br />

tower oscillations is achieved by input-output<br />

pole placement controller. On the other hand<br />

full state feedback controller has shown that is<br />

capable <strong>of</strong> reducing the tower oscillations even<br />

under turbulent conditions. The achieved<br />

damping <strong>of</strong> tower oscillations leads to fatigue<br />

reduction what enables production <strong>of</strong> lighter<br />

and less expensive wind turbines. This can<br />

result in reduction <strong>of</strong> the price for electrical<br />

energy generated by wind turbines.<br />

Acknowledgements<br />

This work was financially supported by Končar –<br />

Electrical Engineering Institute and the Ministry <strong>of</strong><br />

Science Education and Sports <strong>of</strong> the Republic <strong>of</strong><br />

Croatia.<br />

References<br />

[1] T. Burton, D. Sharpe, N. Jenkins, E.<br />

Bossanyi, "<strong>Wind</strong> energy handbook," John<br />

Wiley and sons, 2001.<br />

[2] P. Novak, T. Ekelund, I. Jovik and B.<br />

Schmidtbauer, "Modeling and control <strong>of</strong><br />

variable-speed wind-turbine drive-system<br />

dynamics," Control system magazine, Vol.<br />

15, No. 4, pp. 28-38, 1995.


[3] F. D. Bianchi, H. De Battista and R.J.<br />

Mantz, "<strong>Wind</strong> turbine control system,<br />

principles, modeling and gain scheduling<br />

design," Springer, 2006.<br />

[4] E.L. van der Ho<strong>of</strong>t, P. Schaak and T.G. van<br />

Engelen, "<strong>Wind</strong> <strong>Turbine</strong> Control<br />

Algorithms, Dowec WP1 – task 3 ECN-C—<br />

03-111," ECN <strong>Wind</strong> Energy, Petten, The<br />

Netherlands, 2003.<br />

[5] Bladed Theory manual, GH report,<br />

282/BR/009, 2003.<br />

[6] W.S. Levine (editor), "The control<br />

handbook," The electrical engineering<br />

handbook series, CRC press, 1996.<br />

[7] M. Jelavić, N. Perić and S. Car, "Estimation<br />

<strong>of</strong> wind turbulence model parameters," in<br />

Proc. 2005 International Conference on<br />

Control and Automation, pp. 89-94,<br />

Budapest, Hungary, 2005.<br />

[8] IEC 1400-1, <strong>Wind</strong> turbine generator<br />

systems - Part 1: Safety requirements, Third<br />

edition, 2005.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!