Вы находитесь на странице: 1из 6

Observer-Based Decentralised Control of a Wind

Turbine with a Hydrostatic Transmission


Harald Aschemann and Julia Kersten
Chair of Mechatronics, University of Rostock, Germany
Email: {Harald.Aschemann, Julia.Kersten}@uni-rostock.de

Abstract— In this paper, a decentralised control approach becomes possible [4]. The superior damping characteristics
for a 5 MW wind turbine with a hydrostatic transmission is of a hydraulic transmission, cf. [5], also promises a positive
presented that covers the whole range from low to very high impact on the overall system reliability and lifetime.
wind speeds. In addition to a linear control of the pitch angle,
a multi-variable gain-scheduled PI state feedback control based The aim of this work is to both reduce the implementation
on LQR techniques is proposed for the angular velocities of both effort and to extend earlier work presented in [9]. Instead of a
the rotor and the generator. The sixth-order simulation model central control approach, a decentralised concept is considered
comprises the complete drive train dynamics, the generator in this paper: the pitch angle is controlled by a linear control,
torque dynamics as well as the rotor aerodynamics and is derived whereas a gain-scheduled multi-variable tracking control is de-
from first principles. To reduce the implementation effort, the
multi-variable control structure is based on a simplified state- signed for the angular velocities of the rotor and the generator.
space model with three states and two inputs. Moreover, a The paper is organised as follows: Section II presents the
reduced-order observer estimates the rotor torque as well as an derivation of a sixth-order simulation model for the wind
unknown leakage volume flow for a disturbance compensation. turbine with hydrostatic transmission. Section III points out
The control performance is illustrated by simulation results, how the steady-state desired values for the whole range of
which show an excellent tracking behaviour for the controlled
variables. wind speeds are computed for the controlled outputs. The
control design – based on a simplified system model with
I. I NTRODUCTION neglected actuator dynamics – is discussed in Section IV.
The majority of modern wind turbines in operation today is Taking advantage of extended linearisation techniques, the PI
equipped with pitch actuators. Thereby, the generated power state feedback control law is derived and combined with a
can be kept constant by increasing the pitch angle of the feedforward control. In Section V, a reduced-order disturbance
rotor blades at high wind speeds. Usually, the generator shaft observer is proposed and used for a disturbance compensation.
is mechanically connected to the rotor shaft via a gearbox, Corresponding simulation results are presented and discussed
which is used to step up the angular velocity of the slower in Section VI. Finally, conclusions and an outlook are given
turning rotor shaft. Power electronics are usually employed in Section VII.
to match the frequency of the generated power to that of the
grid by using either synchronous or double-fed asynchronous II. M ATHEMATICAL M ODELLING
machines in combination with current converters. In this paper, the simplest form of a hydrostatic transmission
A hydrostatic transmission represents a type of continuously is regarded: it consists of a hydraulic pump as well as a
variable transmission that uses hydraulic fluid to transmit motor connected in a closed circuit, where only the volumetric
power. Hydrostatic transmissions have been employed suc- displacement of the motor can be varied by a corresponding
cessfully for many years, especially in mobile hydraulics. displacement unit. Thereby, any desired transmission ratio can
The implementation of hydrostatic transmissions in the field be attained within a certain predefined boundary determined by
of wind energy is not a new concept. Already in 1980, an the system parameters. Fig. 1 depicts the drive train structure
early experimental turbine with a hydrostatic transmission of a wind turbine with a hydrostatic transmission. For the
was built. Due to its high hydraulic losses, however, this
attempt turned out to be unsuccessful, see [1] and [2]. Instead, High Pressure
Side
advances in power electronics led the way for the takeover of JR JM
mechanical gear boxes, which are common today. Nowadays, M R , ωR MM , ωM
the application of hydrostatic transmissions for wind turbines
is gathering renewed research interest. This is due to the
fact that a hydrostatic transmission in a wind turbine drive Rotor Side Motor Side
train offers several benefits over the conventional mechanical Low Pressure
Side
gear box. Among these is the potential to increase the annual
energy production by extending the useful wind speed range Fig. 1. Drive train of a wind turbine with a hydrostatic transmission.
of the turbine, cf. [3]. Moreover, an operation at higher
aerodynamic efficiencies over a wider range of wind speeds simulation, a corresponding symbolic system model for the

978-1-4799-8701-6/15/$31.00 ©2015 IEEE. 868


wind turbine is derived from first principles that includes all leakage volume flow of the hydrostatic transmission from the
dominant dynamic effects, such as the rotor aerodynamics, the high pressure to the low pressure side. The unknown volume
drive train dynamics and the actuator dynamics. The drive train flow qU accounts for model uncertainty. The dynamics of the
model has three state variables: the rotor angular velocity ωR , motor displacement unit, which allows for changing the motor
the difference pressure Δp of the hydrostatic transmission, and tilt angle, is characterised by a first-order lag behaviour
the angular velocity ωG of the generator. The corresponding
model structure is illustrated in Fig. 2. The aerodynamic torque 1
α̃˙ M = (uα − α̃M ) , (5)

βd Pitch
Actuator where Tα denotes the time constant. The angular velocity of
β the generator is described by a first-order differential equation
v Aero- uG
dynamics Generator
1
M R ωR M G ωG ω̇G = (ṼM Δp α̃M − dG ωG − MG ) (6)
ωR Hydraulic Δp Hydraulic u JG
Rotor α
Motor
Δp Pump ωG with the combined mass moment of inertia JG of both the
Fig. 2. Flow diagram of the wind turbine with hydrostatic transmission.
motor and the generator, the damping coefficient dG and
the load torque of the generator MG . The generator torque
extracted by the rotor from the wind is given by the following dynamics is governed by
nonlinear expression
1
1 ṀG = (uG − MG ) , (7)
MR = ρ A R CQ (λ, β) v 2 , (1) TG
2
where ρ denotes the density of air, R the rotor radius, A = where TG is the corresponding time constant. A more detailed
Rπ 2 the rotor area, β the blade pitch angle, v the wind speed model of the generator can be found in [15] and is not
and λ the tip-speed ratio according to envisaged in this paper. The torque input is denoted by uG ,
the actual one is given by MG . The dynamics of the pitch
ωR R
λ= . (2) angle axis – representing a fast underlying control loop – can
v be modelled as a first-order lag system
The rotor angular velocity is denoted as ωR . The coefficient
CQ (λ, β) represents the torque coefficient, which is a nonlin- 1
ear function of the tip-speed ratio and the blade pitch angle. In β̇ = (βd − β) , (8)

this work, CQ is given by an analytic expression which stems
from [8]. The power coefficient CP is directly related to the with the time constant Tβ . The desired pitch angle uβ = βd
torque coefficient according to CP = CQ · λ. The equation of serves as input variable, whereas the actual pitch angle is
motion for the rotor is governed by the following first order denoted by β. Due to physical limitations of the pitch actuator,
differential equation the pitch angle as well as the pitch rate are restricted to the
1   ranges 0 ≤ βd ≤ 90 deg and |β̇d | ≤ 10 degs , respectively. The
ω̇R = MR (λ, β) − dR ωR − ṼP Δp , (3) simulation model can be stated by the following sixth-order
JR
nonlinear state-space representation
where JR denotes the overall mass moment of inertia of
the entire rotor as well as the blades. The parameter dR
ẋS = fS (xS , uS , τS ) (9)
characterises the velocity proportional friction of the rotor.
The pressure difference between the high and low pressure
with the state vector xS = [ωR , Δp, ωG , β, MG , α̃M ]T , the
hoses of the hydrostatic transmission is denoted by Δp. The
input vector uS = [βd , uG , uα ]T , and the disturbance vector
volumetric displacement of the pump is given ṼP = V2πP , where
τS = [MR , qU ]T . The system parameter used in the simulation
VP is the volume provided per revolution. The dynamics of the
model are listed in Table I.
difference pressure in the hydraulic transmission is described
by
2 III. M ODEL - BASED D ERIVATION OF THE O PERATING
Δṗ = (ṼP ωR − ṼM ωG α̃M − kl Δp − qU ) , (4) S TRATEGY
CH
where ωG denotes the rotational speed of the generator and The operation range of the turbine is characterised by three
motor. The normalised motor swash plate angle is given by distinct regions with individual control objectives. In general,
αM
α̃M = αM,max and is used to vary the motor volumetric the efficiency should be maximized while simultaneously
displacement ṼM = V2π M
. The hydraulic capacitance CH preventing the turbine from any damage. The regions are
quantifies the energy storage capability of the hydraulic fluid, characterised by the four characteristic points and connecting
whereas the leakage coefficient kl characterises the nominal curves shown in Fig. 3.

869
TABLE I
S YSTEM PARAMETERS C B
0.5
Parameter Symbol Value A
0.4

Cp
kg
Density of air ρ 1.23 m3
0.3
Rotor radius R 63 m
Rotor inertia JR 3.88 · 107 kg · m2 0.2
Rotor damping coefficient dR 10 Nms
rad
0.1
m3
Pump volumetric displacement ṼP 1.50 · 10−3 rad 0
Hydraulic capacity CH 8.84 · 10−10 m2 0
Leakage coefficient (nominal) kl 4.31 · 10−13 m3 5
rad
10 D
Motor volumetric displacement ṼM 0.08 V/A
15 16
12
Generator inertia JG 5.34 · 102 kg · m2 20 8
◦ 4
Nms β in 25 0 λ
Generator damping coefficient dG 10 rad
Generator time constant TG 0.06 s
Fig. 3. Operating points on the CP surface.
Pitch actuator time constant Tβ 0.06 s

TABLE II C. Region III


PARAMETER VALUES AT O PERATING P OINTS
In this region, the power is kept constant at its rated value.
The corresponding curve segment in Fig. 3 is [C, D]. Note that
A B C D
the turbine does not operate at the locally maximum CP value
MR [kNm] 0.10 · 106 1.92 · 106 3.63 · 106 3.63 · 106
but on a curve characterised by MR = const. Therefore, the
MG [kNm] 0.42 · 104 1.50 · 104 4.00 · 104 4.00 · 104 rotor angular velocity remains constant as well. An alternative
ωR [1/s] 0.72 1.00 1.38 1.38 solution is given in [10], where the rotor torque is represented
ωG [1/s] 126 126 126 126 by a monotonically increasing function of the wind speed.
λ 15.2 7.9 7.9 4.3 Given this invertible characteristic, the wind speed can be
v [m/s] 3.0 8.0 11.0 20.0
uniquely determined by the estimated torque provided by an
observer. To maintain the rated power of the turbine constant,
the aerodynamic efficiency must be reduced by varying β
accordingly. Given constant values MR = const and ωR =
A. Region I
const, the pitch angle β and the tip-speed ratio λ can be
Once the rotor has reached its cut-in speed in point A, the determined for each wind speed. The functional dependency
generator starts to produce electrical power. The corresponding of the pitch angle w.r.t. the wind speed can be approximated
operating region is given by the curve segment [A, B] in Fig. 3, by a forth-degree polynomial
where the efficiency attains the maximum possible value. The
β 4 = p 4 · v 4 + p 3 · v 3 + p 2 · v 2 + p 1 · v − p0 , (12)
pitch angle β is fixed to zero in the first two regions and is
only used at very large wind speeds in region III. The rotor with the corresponding coefficients pi , i ∈ {0, .., 4}. By
angular velocity is a linear function of wind speed according increasing β, the operating point moves towards point D,
to hence, reducing the aerodynamic efficiency of the rotor. In
ωR,1 = k1 · v + c1 , (10) the event of reaching point D, the turbine is shut down for
safety reasons.
where the constants k1 and c1 are chosen such that the
operation reaches point B of Fig. 3. D. Calculation of desired values for the states variables
The steady-state values for the controlled outputs are de-
B. Region II picted in Fig. 4 in dependence on the wind speed. These
At the curve segment [B, C] the maximum efficiency with desired values can be determined for each wind speed by a
CP = CP∗ is achieved corresponding to the tip-speed ratio steady-state analysis of the system model (9).
λ = λ∗ . The second region acts as a transition between the First, the steady-state difference pressure Δp̄ is determined
optimal and the rated operation. The rotor angular velocity is from the steady-state rotor torque M̄R as well as the angular
calculated by the linear equation velocity ω̄R of the rotor as follows
M̄R − dR ω̄R
ωR,3 = k3 · v , (11) Δp̄ = . (13)
ṼP
where the constant k3 is chosen such that the rotor reaches Considering a vanishing leakage volume flow qU = 0, the
the rated power at point C, with its rated angular velocity. steady-state motor tilt angle which corresponds to the required

870
1.4 127 Obviously, the input matrix B (Δp̄) is adapted to the differ-
ence pressure Δp̄ corresponding to the actual wind speed as

ωG in [s−1 ]
ωR in [s−1 ]

1.2
depicted in Fig. 4. The corresponding output equation becomes
1 126 ⎡ ⎤
    ωR
0.8 ωR 1 0 0 ⎢ ⎥
ωR ωG
ωG
= ⎣Δp⎦ . (16)
0.6
4 6 8 10 12 14 16 18
125
20
0 0 1
    G ω
y C  
20 x
15 In the sequel, the dependency on the difference pressure is
β in [◦ ]

10 omitted for simplicity. To design a PI state feedback, the model


5
is augmented with two new states representing the integrated
tracking errors regarding the controlled variables
0
4 6 8 10 12 14 16 18 20
   t 
v in [m/s] ωR,I (ωR,d (v̂) − ωR ) dτ
xI = = 0
t . (17)
ωG,I
0
(ωG,d − ωG ) dτ
Fig. 4. Desired values in dependence on the wind speed with Point A at the
wind speed of 3 ms
, B at 8 ms
, C at 11 m
s
and D at 20 ms
. The subscript d denotes desired values, whereas I charac-
terises the two integrator output variables. The desired value
ωR,d is a filtered version of the corresponding steady-state
difference pressure is calculated as value discussed in section III. Note that ωG,d remains con-
 
ṼP ω̄R kl M̄R − dR ω̄R stant over the entire operating range. The augmented system
¯
α̃m = − . (14)
ṼM ω̄G ṼP ṼM ω̄G description becomes
          
Finally, the steady-state generator torque ẋ A 0 x B E 0
= + u+ τ+ w , (18)
¯ M − dG ω̄G
M̄G = ṼM Δp̄ α̃ (15) ẋI −C 0 xI 0 0 I
    
can be determined, which is required to maintain a constant Aa xa Ba
angular velocity of the generator, ω̄G = const. where the subscript a denotes the augmented system, and w =
 T
IV. C ONTROL D ESIGN BASED ON E XTENDED ωR,d ωG,d the vector of desired values. The output matrix
L INEARISATION is augmented similarly as Ca = C 0 . A time-weighted
A popular approach to wind turbine control considers sep- quadratic cost function is introduced as follows

arate controllers for each operating region, see [12] and [13]. 1 ∞ T 
In this paper, a decentralised control structure is designed J= xa Qa xa + uT Ra u e2αt dt , (19)
2 0
for all regions from small up to very large wind speeds,
where Qa and Ra are positive semi-definite and positive
cf. [9] and [16]. The decentralised control structure involves
definite weight matrices, respectively. Here, the coefficient
an underlying control of the pitch angle, given by (8), and a
α > 0 guarantees a stability margin in the LQR design. The
multi-variable control of the drive train considering the motor
LQR feedback gain matrix Ka = K + KP C −KI that
tilt angle uα = α̃M and the generator torque uG = MG
minimises the quadratic cost function in (19) is computed by
as ideal input variables. The multi-variable control design is,
solving the algebraic Riccati equation online with measured
hence, based on a simplified reduced-order model, where the
values for the difference pressure Δp. Here, the gain matrix
asymptotically stable dynamics for the motor displacement
KP can be chosen as the identity matrix KP = I. As an
unit and the generator torque are neglected. The nonlinear
alternative to an online solution, the Riccati equation could
system representation based on the remaining differential
be solved offline for a reasonable number of set points in the
equations is written in quasi-linear form with state-dependent
relevant operating range. Using look-up tables, a gain schedul-
matrices, cf. [17]. This leads to
⎡ ⎤ ⎡ dR ⎤⎡ ⎤ ing of the control gains becomes possible. By including a
ω̇R − JR − ṼJRP
0 ωR feedforward control law uv = Kv0 w, the system is expressed
⎢ ⎥ ⎢ ⎢
⎥⎢ ⎥
⎥ as
⎣Δṗ⎦ = ⎣ CHP − CHl
2 Ṽ 2k
0 ⎦ ⎣Δp⎦     
ω̇G 0 0 − JdG ωG ẋ A − B K − B K P C B KI x
   G
   =
ẋ x ẋI −C 0 xI
⎡ ⎤ ⎡
A
⎤  
0 0   1
0   Ag
⎢ 1 ⎥ JR   (20)
MG ⎢ ⎥ MR
+⎢
⎣ JG − 2ṼCMHω̄G ⎥
⎦ α̃M + ⎣ 0 − C2H ⎦ B KP
qU + Ba Kv0 w + w.
  I
0 ṼM Δp̄
  0 0  

J
 G  u   τ
uv  
E Bw,g
B(Δp̄)

871
The feedforward gain matrix Kv0 is derived by fulfilling the H = diag(h11 , h22 ) is the observer gain matrix. The state
following design condition equation for the internal state vector z is given by
−1
I = Ca (−Ag ) (Ba Kv0 + Bw,g ) . (21) ẑ˙ = Φ (xO , α̃M , τ̂ ) . (26)

The four unknowns in Kv0 are chosen in such a way the right The observer gain H and the vector Φ have to be chosen in
hand side becomes a unity matrix. This guarantees the tracking such a way that the steady-state observer error τ̃ = τ − τ̂
of ramps without producing any subsequent errors. A static converges to zero. Thus, Φ can be determined by demanding
compensation using the disturbance estimate τ̂ = [M̂R , q̂U ]T a vanishing steady-state estimation error according to
is calculated according to the following design condition τ̃˙ = 0 = τ̇ − H ẋO − Φ (xO , α̃M , τ̂ − 0) . (27)
−1 −1
0 = C (s I − AR ) B Kdc0 + C (s I − AR ) E, (22) Considering τ̇ = 0, (27) yields
with AR = A − B K. The overall multi-variable control law Φ (xO , α̃M , τ ) = −H fO (xO , α̃M , τ ) . (28)
comprises feedforward and feedback control action as well as
a disturbance compensation. It is implemented as The linearised error dynamics must be asymptotically stable.
⎡ ⎤ Thus, all eigenvalues of the Jacobian are placed in the left
x complex half-plane according to
 ⎢ ⎥  
u = − K −KI −KP ⎣xI ⎦ + Kv0 w + Kdc0 τ̂ , (23) ∂Φ (xO , α̃M , τ ) ! −sB1 0
ẋI = , (29)
∂τ 0 −sB2
see Fig. 5. During the simulation, the gain matrices are with positive values sB1 > 0, sB2 > 0, where the observer
adapted to the difference pressure Δp̄ that corresponds to the gains follow directly from the comparison (29). The benefits
wind speed v. An anemometer situated on top of the nacelle of the observer-based control approach become obvious in the
following section.

 VI. S IMULATION R ESULTS
  In the sequel, simulation results are presented for the
 
 



 proposed decentralised control structure. For this purpose, a


 
  
   sequence of several changes regarding the wind speed are

  
   specified that covers the whole operation range from the cut-in
wind speed up to very high wind speeds. Moreover, a realistic

noise process is considering in the simulation. The wind profile
 as well as the simulated difference pressure Δp are depicted
 
  in Fig. 6. The control gains are adapted to these varying

 
difference pressure by solving the algebraic Riccati equation
online with a reduced rate of 1 Hz. Each of these wind speed
Fig. 5. Control implementation with disturbance observer.
changes result in attaining a different operating region as
discussed in Section III. The corresponding simulated values
measures the wind speed, as presented in [14]. for the controlled variables are plotted in Fig. 7 along with
their desired values to allow for a comparison. It becomes
V. R EDUCED - ORDER DISTURBANCE OBSERVER
obvious that the proposed control structure leads to very small
For the implementation of the proposed control approach, a tracking errors for the pitch angle on the one hand, and for
reduced-order disturbance observer is introduced. Considering the angular velocities of the rotor and the generator on the
all state variables as measurable, estimates are required only other hand. Fig. 8 shows simulation results of the reduced-
for the unknown system disturbances – the rotor torque MR order observer by comparing estimated and simulated values
and the leakage volume flow qU . Therefore, the first two state of the disturbance MR and the leakage qU , which match well.
equations The almost constant estimate M̂R in region III indicates that
   −dR ωR −ṼP Δp+MR
 the overall power remains constant as well.
ω̇R
ẋO = JR
= 2(ṼP ωR −ṼM ωG α̃M −kl Δp−qU ) (24) VII. C ONCLUSION AND O UTLOOK
Δṗ
CH
In this paper, a sixth-order state-space model of a 5 MW
are extended with two integrators as disturbance models wind turbine equipped with a hydrostatic transmission has
ẋO = fO (xO , α̃M , τ ) , τ̇ = 0 , (25) been derived. The control structure involves two separate
decentralised controllers: a linear control of the pitch angle,
where xO contains the measurable state variables ωR and and a multi-variable control of the angular velocities of the
Δp. The disturbance vector is given by τ = [MR , qU ]T . The rotor and the generator. By applying a model-order reduction
estimated disturbances follow from τ = HxO + z, where as well as extended linearisation techniques, a third-order

872
18 model for the multi-variable control design can be stated
16 in quasi-linear form with state-dependent matrices. A gain-
14
s ]

scheduling tracking control scheme has been implemented


v in [ m

12
10
8 using an optimal proportional-integral state feedback control
6 in combination with feedforward control. Moreover, the rotor
4
2 torque as well as an unknown leakage flow are estimated
0 50 100 150 200 250 300 350 400
by a reduced-order observer and employed for a disturbance
400 compensation. The proposed control structure has been inves-
Δp in [bar]

300 tigated by simulations, and has shown an excellent tracking


200
100
performance within its entire operating range. Future work
0 will address an active damping of tower oscillations.
-100
-200 R EFERENCES
0 50 100 150 200 250 300 350 400
[1] V. Nelson. Wind Energy: Renewable Energy and the Environment. Boca
t in [s]
Raton, FL: CRC Press, 2009, p. 125.
[2] W. Rampen. ”Gearless Transmissions for Large Wind Turbines:
Fig. 6. Measured wind speed v and the resulting difference pressure Δp
The History and Future of Hydraulic Drives.” Internet:
serving as scheduling variable.
www.artemisip.com/Pictures/GearlesstransmissionsBremenNov06.pdf,
[Feb. 28, 2012].
[3] C. Gorla and P. Cesana. ”Efficiency Models of Wind Turbines Gearboxes
1.4
ωR in [s−1 ]

des
sim with Hydrostatic CVT.” Balkan Journal of Mechanical Transmissions
1.2 (BJMT), vol. 1, pp. 17-24, 2011.
1 [4] ”Advanced Wind Turbine Drivetrain Concepts: Workshop Report”, Inter-
0.8 net: www.nrel.gov/docs/fy11osti/50043.pdf, [Feb. 28, 2012].
[5] N.F.B. Diepeveen and A.J. Laguna. ”Dynamic Modeling of Fluid Power
0 50 100 150 200 250 300 350 400
Transmissions for Wind Turbines,” presented at EWEA Offshore, Ams-
127 terdam, Netherlands, 2011.
ωG in [s−1 ]

des
126.5 sim [6] L. Mangliardi and G. Mantriota. ”Automatically Regulated C.V.T. in Wind
126 Power Systems.” Renewable Energy, vol. 4, pp. 229-310, Oct. 1993.
125.5 [7] J. Jonkman, S. Butterfield, W. Musial and G. Scott. ”Definition of a 5-MW
125 Reference Wind Turbine for Offshore System Development.” Internet:
0 50 100 150 200 250 300 350 400 homes.civil.aau.dk/rrp/BM/BM8/k.pdf, Feb. 2009 [Jun. 17, 2011].
15
des
[8] S. Georg, H. Schulte and H. Aschemann. ”Control-Oriented Modelling of
10
sim Wind Turbines using a Takagi-Sukeno Model Structure,” Proc. of FUZZ-
β in [◦ ]

IEEE, Brisbane, Australia, 2012.


5 [9] B. Dolan and H. Aschemann. ”Control of a Wind Turbine with a
0
Hydrostatic Transmission - an Extended Linearisation Approach,” Proc.
0 50 100 150 200 250 300 350 400 of MMAR, Miedzyzdroje, Poland, 2012.
t in [s] [10] H. Aschemann and J. Kersten. ”Active Tower Damping for an Innovative
Wind Turbine with a Hydrstatic Transmission,” under review for Proc. of
IECON, Yokohama, Japan, 2015.
Fig. 7. Comparison of desired and simulated values for the controlled
[11] T. Burton, N. Jenkins, D. Sharpe and E. Bossanyi. Wind Energy
variables.
Handbook. Chichester, UK: Wiley, 2011, p. 485.
[12] K.A. Stol, W. Zhao and A.D. Wright. ”Individual Blade Pitch Control
x 106 for the Controls Advanced Research Turbine.” Journal of Solar Energy
4
Engineering, vol. 128, pp. 498-505, Nov. 2006.
MR in [kNm]

3 [13] R. Burkart, K. Margellos and J. Lygeros. ”Nonlinear Control of Wind


Turbines: An Approach Based on Switched Linear Systems and Feedback
2 Linearisation,” in Proc. 50th IEEE Conference on Decision and Control
and European Control Conference (CDC-ECC), 2011, pp. 5485-5490.
1 M̂R
MR
[14] S. Heier. ”Windkraftanlagen - Systemauslegung, Netzintegration und
0 Regelung (in German),” 5th edition, 2009, Vieweg+Teubner, pp. 387-396.
0 50 100 150 200 250 300 350 400 [15] J. Fortmann. ”Modeling of Wind Turbines with Doubly Fed Generator
−6 Systems,” 1st edition, 2015, Springer Vieweg, pp. 81-111.
x 10
20 [16] C.L. Bottasso and A. Croce. ”Power Curve Tracking with Tip Speed
Constraint using LQR Regulators.” Internet: www.aero.polimi.it/
15
˜bottasso/downloads/CLBottasso\_ACroce\_2009c.pdf,
qU in [ ms ]
3

10 Mar. 2009 [Mar. 6, 2012].


5 [17] B. Friedland. ”Advanced Control System Design,” 1996, Prentice Hall.
q̂U
0 qU
-5
0 50 100 150 200 250 300 350 400
t in [s]

Fig. 8. Comparison of estimated and simulated values of the rotor torque


MR and the leakage volume flow qU .

873

Вам также может понравиться