You are on page 1of 12

Engineering Applications of Artificial Intelligence 34 (2014) 1–12

Contents lists available at ScienceDirect

Engineering Applications of Artificial Intelligence


journal homepage: www.elsevier.com/locate/engappai

Active sensor fault tolerant output feedback tracking control for wind
turbine systems via T–S model
Montadher Sami Shaker a,n, Ron J. Patton b
a
Department of Electrical Engineering, University of Technology, Baghdad, Iraq
b
School of Engineering, University of Hull, Hull HU6 7RX, UK

art ic l e i nf o a b s t r a c t

Article history: This paper presents a new approach to active sensor fault tolerant tracking control (FTTC) for offshore
Received 17 September 2013 wind turbine (OWT) described via Takagi–Sugeno (T–S) multiple models. The FTTC strategy is designed
Received in revised form in such way that aims to maintain nominal wind turbine controller without any change in both fault and
11 February 2014
fault-free cases. This is achieved by inserting T–S proportional state estimators augmented with
Accepted 14 April 2014
proportional and integral feedback (PPI) fault estimators to be capable to estimate different generators
Available online 28 May 2014
and rotor speed sensors fault for compensation purposes. Due to the dependency of the FTTC strategy on
Keywords: the fault estimation the designed observer has the capability to estimate a wide range of time varying
Active fault tolerant control fault signals. Moreover, the robustness of the observer against the difference between the anemometer
Fault estimation
wind speed measurement and the immeasurable effective wind speed signal has been taken into
Tracking control
account. The corrected measurements fed to a T–S fuzzy dynamic output feedback controller (TSDOFC)
T–S fuzzy systems
Dynamic output feedback control designed to track the desired trajectory. The stability proof with H1 performance and D-stability
Wind turbine control constraints is formulated as a Linear Matrix Inequality (LMI) problem. The strategy is illustrated using a
non-linear benchmark system model of a wind turbine offered within a competition led by the
companies Mathworks and KK-Electronic.
& 2014 Elsevier Ltd. All rights reserved.

1. Introduction significantly wide range of weather conditions and minimise both


manufacturing and maintenance costs. To maximise the amount of
Owing to inherent limitations in different kinds of the well- the annual power production an increase in wind turbine size has
known fossil fuel and nuclear energy sources, e.g. carbon footprint, been suggested, another opportunity is the development of a
rapidly increasing fuel prices or probability of catastrophic effects variable speed wind turbine which enhances both quality and
of nuclear station malfunction, the last two decades have wit- the amount of the power production compared with the fixed
nessed a rapid growth in the use of wind energy. Although, it is speed wind turbines (Bianchi et al., 2007). Furthermore, to reduce
considered a promising source of energy, depending on naturally the effects of obstacles and roughness of terrain that increase
generated wind forces, there are several very significant challenges wind force turbulence, OWTs are currently being developed and
to efficient wind energy conversion for electrical power transfor- installed (Stewart and Lackner, 2013).
mation (Marden et al., 2013; Odgaard et al., 2013). In the last years, a number of publications consider the effect of
Wind turbine systems demand a high degree of reliability and wind turbines malfunction is noticed with focusing on designing
availability (sustainability) and at the same time are characterised fault tolerant control (FTC) (Sloth et al., 2011; Kamal et al., 2012;
by expensive and safety critical maintenance work (Verbruggen, Sami and Patton, 2012b, 2012a; Badihi et al., 2013; Jain et al., 2013;
2003). The recently developed OWTs are foremost examples Odgaard et al., 2013; Simani and Castaldi, 2013, 2014) and fault
since OWT site accessibility and system availability are not always monitoring systems (Amirat et al., 2009; Hameed et al., 2009; Wei
ensured during or soon after malfunctions, primarily due to and Liu, 2010; Johnson and Fleming, 2011; Djurovic et al., 2012). In
changing weather conditions. Hence, the main challenges for the Sloth et al. (2011), the authors proposed linear parameter-varying
deployment of wind turbine systems are to maximise the amount FTC systems for pitch actuator faults occurring in the full load
of good quality electrical power extracted from wind energy over a operation. In Kamal et al. (2012) a T–S fuzzy observer-based FTC
design is proposed to achieve maximisation of the power extrac-
tion in the presence of generator sensor fault without taking into
n
Corresponding author.
account the trade-off between the tracking of the optimal signal
E-mail addresses: m.s.shaker@uotechnology.edu.iq (M.S. Shaker), and the loads induced in the drive train due to exact tracking of
r.j.patton@hull.ac.uk (R.J. Patton). optimal signal. Jain et al. (2013) present a real-time projection-

http://dx.doi.org/10.1016/j.engappai.2014.04.005
0952-1976/& 2014 Elsevier Ltd. All rights reserved.
2 M.S. Shaker, R.J. Patton / Engineering Applications of Artificial Intelligence 34 (2014) 1–12

based approach to tolerate faults occurring in a wind turbine output to follow the reference signals and hence the control signal
system. In Sami and Patton (2012b, 2012a) the robustness of will no longer be suitable for the system under control. The
sliding mode control have been utilised to design FTC for OWT. controller reconfiguration approach can handle more general
Badihi et al. (2013)) present an approach for designing a fault faults and/or failure cases through either off-line or on-line
detection and diagnosis (FDD) and FTC scheme for a wind turbine variation of the structure and/or the parameters of the controller
using fuzzy modelling and control. Simani and Castaldi (2014) based on the information delivered from an FDD unit. However,
proposed a fault tolerant controller to accommodate actuator fault the main challenge of this method is that the time required to
using the on-line fault estimate signal generated by an adaptive reconfigure the control system must be as low as possible. In fact,
filter. The fuzzy approach to wind turbine controller design in the this is very important in practice where the time windows during
presence of modelling and measurement errors has also been which the system remains stabilisable in the presence of a fault
proposed in Simani and Castaldi (2013). are very short. This is especially the case for unstable open-loop
Generally, wind turbines have non-linear aerodynamics and systems, e.g. the unstable double inverted pendulum example
this limits the use of a linear systems approach without due care to (Niemann and Stoustrup, 2005). On the other hand, as an exten-
robustness issues. Furthermore, wind turbines have a stochastic sion to the use of FDD base control reconfiguration, the estimation
and uncontrollable driving force as input in the form of wind and compensation approach to FTC is based on the computation of
speed. This, together with overall system nonlinearity limits the fault estimates and a mechanism to compensate these fault effects
ability of linear control strategies to satisfy the control objectives. by the addition of a new compensating signal to the nominal
Hence, an increase in interest in controlling wind turbines through control input or faulty measured output. Clearly, this approach
nonlinear control methods has been noticed in the last years to obviates the need for residual evaluation and parameter identifi-
handle the nonlinearity of the turbine aerodynamics. This is cation steps that are required for FDD based FTC and hence
achieved either through the use of nonlinear models directly in requires no time consuming algorithms for maintaining the
the design (Sami and Patton, 2012b, 2012a) or through the use of performance of the nominal system control law. Moreover, the
multiple-model approaches (Sloth et al., 2011; Kamal et al., 2012; inability of the adaptive control-based FTC to tolerate sensor faults
Badihi et al., 2013). means that fault estimation and compensation represents the all
This paper focuses on the design of FTTC dedicated to optimise round most appropriate method for the sensor fault case of FTC.
the captured wind energy. Based on the wind turbine T–S fuzzy Within the framework of estimation and compensation, this
model, a TSDOFC is designed to achieve the required tracking of section presents the structure of the proposed FTTC strategy for
the optimal rotor speed (ωr opt ) and the reduction of loads OWT control problems based on robust fault estimation and
affecting the drive-train through minimising the L2 norm of the compensation of rotor rotational speed sensor faults f^ sr and/or
wind variation on the torsion angle of the drive train. The FTTC is generator rotational speed sensor faults f^ sg whilst maintaining the
achieved by hiding the effects of different generators and rotor performance and stability of the nominal control system during
sensors fault from the controller inputs through the insertion of both faulty and fault-free cases. It is clear from the architecture
the PPI observer (PPIO) capable of handling the case of time- shown in Fig. 1, the proposed FTTC scheme is based on the
varying fault signals. The proposed strategy overcomes the depen- combination of (a) robust TSDOFC and (b) estimates of the f^ sr
dence and limitation of requiring full state measurements. The and/or f^ sg via the T–S fuzzy PPIO. This strategy can be considered
technique also focuses on minimising the difficulties of designing as a “fault-hiding” approach to FTC where the main aim of fault-
observer based control systems that require special pole place- hiding is to maintain the same controller in both faulty and fault-
ment conditions since these conditions cannot hold perfectly in free system cases. Specifically, the T–S fuzzy PPIO will hide sensor
the multiple model framework due to global stability constraints. faults through fault estimation and compensation so that the
The design also locates the controller and observer poles of each TSDOFC always receives the fault free measurements.
T–S model system to lie within a disc region of the complex plane. It should be noted that the T–S fuzzy model consists of local
Finally, the design is formulated as a linear matrix inequality linear input–output relations of the nonlinear system. The overall
problem that can be solved easily using MATLAB software. The fuzzy model is achieved by connecting the local linear models by
simulation results are based using a non-linear benchmark system membership functions yielding the global model of the system. It
model of a wind turbine offered within a competition led by the is important to note that a T–S fuzzy controller is a model-based
companies Mathworks and KK-Electronic (Odgaard et al., 2009, control approach. The procedure of designing T–S fuzzy controller
2013 ). and/or T–S fuzzy observer should start from deriving the T–S fuzzy
model. Hence, the next section presents the nonlinear model of
the wind turbine and the derivation of its corresponding T–
2. The proposed active FTTC S model.

Generally, active FTC (AFTC) systems are designed to handle the


occurrence of system faults on-line by using fault estimation/
compensation methods, adaptive control or controller reconfi-
guration mechanisms. All AFTC methods involve some advantages
and disadvantages. The operation philosophy of adaptive control
fits very well with the AFTC approach. This is due to the ability of
adaptive control systems to adjust controller parameters on-line
based on measured signals. Clearly, the use of adaptive control
methods as an approach to AFTC obviates the need for fault
detection and diagnosis (FDD) unit. However, in this method,
sensor faults represent the most challenging fault scenario for
AFTC and have rarely been considered in fault tolerant adaptive
control methods. For example, output feedback adaptive tracking
control can tolerate actuator and/or system faults, whereas, if
sensor faults have occurred the adaptation will force the faulty Fig. 1. Active sensor FTTC scheme.
M.S. Shaker, R.J. Patton / Engineering Applications of Artificial Intelligence 34 (2014) 1–12 3

3. Wind turbine modelling state space model of wind turbine is given as


)
x_ ¼ AxðtÞ þ Bu þ Ev
A wind turbine model is obtained by combining the constituent ð7Þ
y ¼ CxðtÞ
component subsystem models that together make up the overall
wind turbine dynamics. The aerodynamic torque (Ta) represents where
the source of nonlinear nature of wind turbine which in turn 2 3
depends on the rotor speed ωr, the blade pitch angle β and the  τ1g 0 0 0 0 0
6 7
effective rotor wind speed v. The aerodynamic power captured by 6 0 0 I 0 0 0 7
6 7
6 7
the rotor is given by 6 0  ω2n I  2ζωn I 0 0 0 7
6 7
P cap ¼ 12 ρπ R C p ðλ; βÞν
2 3
ð1Þ A¼66 0
1 ∂T a
0  ðBdt Jþ Br Þ þ J1 ∂∂Tωar Bdt K dt 7
J 7
J r ∂β ng J r
6 r r r 7
6 1 Bdt ðBdt þ ng Bg Þ K dt 7
where ρ is the air density, R is the radius of the rotor, and Cp is the 6 J 0 0  7
6 g ng J g n2g J g ng J g 7
4 5
power coefficient that depends on the β and the tip-speed-ratio 0 0 0 1  n1g 0
(λ) (TSR) which defined as
2 3 2 3
ωr R 1
0 0
λ¼ ð2Þ 6
τg
7 6 7 2 3
ν 60 0 7 6 0 7 1 0 0 0 0 0
6 7 6 7
6 2 7 6 0 7 60 07
The aerodynamic torque is 60 ωn I 7 6 7 6 1 0 0 0 7
B¼6 7; E ¼ 6 1 ∂T a 7; C ¼ 6 7;
60 0 7 6 J ∂v 7 40 0 0 1 0 05
T a ¼ 12 ρπ R3 C q ðλ; β Þν2 ð3Þ 6 7 6 r 7
6 7 6 0 7 0 0 0 0 1 0
40 0 5 4 5
where C q ¼ ðC p =λÞ is the torque coefficient. 0 0 0
2 3
The drive train is responsible for gearing up the rotor rotational Tg
speed to a higher generator rotational speed. The drive train model 6 β 7
6 7
includes low and high speed shafts linked together by a gear- 6 7 " #
6 β_ 7 T gr
box modelled as a gear ratio. The state space model of the wind 6 7
x¼6 7; u ¼
6 ωr 7 βr
turbine drive train in the form: 6 7
6ω 7
2 3 4 g5
2 _ 3  ðBls Jþ Br Þ Bls
 KJ ls 2 3 21 3
θΔ
ωr 6 r ng J r r 7 ωr Jr 0 " #
6ω 7 6 7 6 7 Ta
4 _g5¼6
Bls ðB þ n B Þ
 ls n2 J g g 76 7 6
ng J g 74 ωg 5 þ 4 0
K ls
 J1 7
6 ng J g g g g 5 Tg ∂T a ∂C
4 5 θ 1
θ_ Δ 1  n1g 0 Δ 0 0 ¼ ρAν3 p
∂β 2ωr ∂β
ð4Þ
∂T a 1 ∂C 1
¼ ρAν3 p  2 ρAν3 C p
where Jr is the rotor inertia, Br is the rotor external damping, Jg is ∂ωr 2ωr ∂ωr 2ωr
the generator inertia, ωg and Tg are the generator speed and
∂T a 1 ∂C 3
torque, Bg is the generator external damping, ng is the gearbox ¼ ρAν3 p þ ρAν2 C p
ratio, and θΔ is the torsion angle. ∂v 2ωr ∂v 2ωr
The electrical system in the wind turbine and the electrical It is clear from the state space model given in Eq. (7) that the
system controllers is much faster than the frequency range used in system matrix A and the disturbance matrix E are not fixed
the benchmark model. Hence, the electrical system is given by the matrices and depend on state variables, the uncontrollable input
following linear relation: v, and the partial derivatives of the usually non-analytical function
1 1 of λ and β, Cp. Hence, to cope with system nonlinearity, a nonlinear
T_ g ¼  T g þ T gr ð5Þ control strategy is required to achieve the aim and objectives of
τg τg
wind turbine operation.
where Tgr is the reference generator torque signal and τg is the
time constant. Finally, the hydraulic pitch system is modelled as a 3.2. The T–S model of the wind turbine
closed-loop transfer function between the measured pitch angle β
and its reference βr: Due to the nonlinear behaviour of the aerodynamic subsystem
and its dependence on the wind speed, it is decided to use the T–S
ω2n
β¼ βr ð6Þ fuzzy model based control strategy to design active sensor FTTC.
s2 þ2ζωn s þ ω2n
Several reasons lead to satisfaction that a T–S fuzzy nonlinear
where ζ is the damping factor and ωn is the natural frequency. For control can cope with wind turbine control requirements, which
the three blade wind turbine system, a transfer function is are
associated with each of the three pitch systems. In cases of no
fault the damping factors are assumed equal.  The T–S fuzzy control makes use of a linear control strategy
locally to produce a nonlinear controller through fuzzy infer-
ence modelling, in terms of fuzzy multiple-modelling.
3.1. The state space model  By increasing the number of premise variables, the T–S fuzzy
model can cover a wider range of operation scenarios which
For controller design purposes the state space model of wind cannot be considered with a linear robust controller. For
turbine is presented in this subsection. The nonlinear model of a example, a linear robust controller is designed based on the
wind turbine is established by combining the individual systems linearised model derived at a specific operating point belong to
given in Eqs. (3)–(6). However, it is clear that the main source of the ideal operation curve given in Fig. 4. Hence, all other
nonlinearity is the aerodynamic subsystem which is usually line- operating regions are considered as modelling uncertainty.
arised in order to predict its effects on all model states. Hence, the Moreover, this controller design always degrades the nominal
4 M.S. Shaker, R.J. Patton / Engineering Applications of Artificial Intelligence 34 (2014) 1–12

required performance in order to take good care of the 3.3. Wind turbine control
modelling uncertainty. On the other hand, by considering the
wind speed and the rotor speed as premise variables in the low In order to best understand the wind turbine control chal-
wind speed range (Region 2 in Fig. 4), the T–S fuzzy model can lenges, the fundamental theory of the wind power extraction
approximate the wind turbine model not only during its ideal process and the upper bound of conversion efficiency C p max ðλ; βÞ
operation curve but it can additionally cover the operation of the captured power P cap to the wind power P wind must first be
scenarios in which the system inputs and outputs deviated clarified.
from ideal operation trajectory. This scenario usually happens Basically, actuator disc theory is used to derive the P cap given in
during wind turbine operation, specifically for large inertia Eq. (1) and the maximum C p ðλ; βÞ. The actuator disc is a generic
wind turbines, since the variation of wind speed is faster than device that has the ability to extract wind energy when it is
rotor speed variations. immersed in airflow passing through a virtual tube (see Fig. 2).
The maximum power captured by disc actuator will be
(Odgaard et al., 2013)
The structure of the nonlinear state space model given in Eq.
(7) is characterised by its common input (B) and common output P cap ¼ 12 ρAν316
27 ¼ 0:59P wind ð10Þ
(C) matrices. This fact plays a vital role in simplification and
Clearly, the three blade variable speed variable pitch wind
conservatism reduction of T–S fuzzy controller design (Tanaka
turbines are a special case of the actuator disc. In this special
and Wang, 2001). For example, the quadratic parameterisation of
structure of actuator disc the blade pitch angles (β), wind speed
the dynamic output feedback controller can be reduced to a linear
(v) and the rotor rotational speed (ωr) are the main variables that
parameterisation dynamic output feedback controller to control
affect the amount of the power captured (i.e. Cp) (see Fig. 3).
the wind turbine.
As wind turbines are driven by a naturally generated wind
The aim is to develop a controller whose gain varies with wind
force, the operating range is divided into four regions according to
speed. For example, in the low wind speed range of operation the
wind speed (see Fig. 4). Region 1, this region is also called the cut-
control aim is to maximise the amount of power extracted from
in region, in which the wind speed is not sufficient to overcome
the available wind power through tracking the optimal rotor
the wind turbine inertia and hence there is no electrical power
rotational speed (ωr opt ) reference signal. Hence, to derive the
generated. Region 2 in which the wind speed is above the cut-in
T–S model with minimum uncertainty, the wind speed (v) and the
and below the rated wind speed, the wind turbine objective here
rotor speed (ωr) are considered as premise variables.
is to maximise the amount of the power harvested from the wind
During the low wind speed of operation (Region 2) the v varies
and transfer it to electrical power. Region 3 in which the wind
within the operating range:
speed is high and rotational speed is equal or above the rated
v A ½vmin ; vmax  m s  1 speed and below the cut-out speed, the objective is to regulate the
generated electrical power to be equal to the rate power. Region
where in the benchmark wind turbine considered in this paper 4 in which the wind speed goes above the upper limit of the
vmin ¼ 4 m s  1 and vmax ¼ 12:5 m s  1 . According to these limits predefined working range.
the other premise variable (ωr) is bounded by Fig. 4 shows that wind turbine operation requires different
control objectives for different ranges of wind speed. Clearly, to
ωr A ½ωmin ; ωmax rad s  1
where ωmin ¼ 0:56 rad s  1 and ωmax ¼ 1:74 rad s  1 . The bounds of
ωr are determined using Eq. (2) using λopt ¼ 8.
The membership function is selected as follows:
9
M 1 ¼ ωωmax
r  ωmin
 ωmin >
>
>
> Fig. 2. Actuator disc.
M2 ¼ 1  M1 =
v  vmin ð8Þ
N 1 ¼ vmax  vmin >>
>
>
N 2 ¼ 1 N 1 ;

Based on these two premise variables four local linear models


of the wind turbine can be determined to approximate the non-
linear system at different operating points in the low range of
wind speed. Hence, Eq. (9) gives the four rule T–S fuzzy model of Fig. 3. Wind turbine power extraction.
the nonlinear wind turbine in Eq. (7):
r
9
>
x_ ¼ ∑ hi ðv; ωr Þ½Ai xðtÞ þ Bu þ Ei v = 8
x 106
i¼1 ð9Þ
>
; 7
Region 1 Region 2 Region 3 Region 4
y ¼ CxðtÞ
6

where h1 ¼ M 1 nN 1 , h2 ¼ M 1 nN 2 , h3 ¼ M 2 nN1 , and h4 ¼ M 2 nN 2 .


Power (watt)

5
4

Remark 1. Clearly, more fuzzy rules can approximate nonlinear 3

systems more accurately. However, as the number of fuzzy rules 2 Total wind power
increases, computation burden and design conservatism are una- 1 Power harvisted by ideal actuator disc
Ideal power harvisted by wind turbine
voidable so that there might be no feasible solution to the set of 0
0 5 10 15 20 25
LMI design constraints. Therefore, the two conditions are partially
Wind speed (m/s)
conflicting with each other, and a trade-off between the number of
fuzzy rules and design conservatism should be made. Fig. 4. Wind turbine region of operation.
M.S. Shaker, R.J. Patton / Engineering Applications of Artificial Intelligence 34 (2014) 1–12 5

Power conversion efficiency by the previous blade. In this case the rotor acts as a rigid
0.5
obstacle and prevents the undisturbed wind from passing
0.4
Low conversion Low conversion
efficiency efficiency through the rotor.
c. t b  t w : This scenario corresponds to the optimal operation in
0.3
which the blade harvests the power from the re-established
3
wind component.
Cp

0.2
λ 0
opt
0.1

0
Hence, wind turbines must be properly controlled to operate in
-2 the vicinity of λopt , at which t b  t w , in order to extract as much
-0.1 wind power as possible.
2 4 6 8 10 12 14

λ
Remark 2.
Fig. 5. Variation of Cp with respect to λ at various β.

 The controller designer must consider several factors that char-


achieve wind turbine control objectives, different system variables
acterise the wind turbine systems such as the requirement of
should be controlled for each specific range of wind speed. The
different control objectives for different ranges of wind speed, the
controlled variables could be investigated from the relationship
nonlinearity of the aerodynamic subsystem, the lack of accurate
between the power conversion efficiency, the wind speed, and the
measurement of the wind speed, and the probability of fault
tip speed ratio as shown in Fig. 5.
occurrence.
Usually, wind speed ν and rotor rotational speed ωr are given in  Exact tracking leads to increase loading on the two drive train
terms of the tip speed ratio λ. The variation of power conversion
shafts and hence can shorten the drive train life time. This also
efficiency with respect to λ and β is given either as a mathematical
produces a highly fluctuating output power, and may even
polynomial or as look-up table. Fig. 5 below shows this relation-
produce a varying direction reference torque signal that can
ship for the benchmark wind turbine considered in this paper.
lead to abnormal generator operation. It is thus very clear that
From Fig. 5 two points must be highlighted:
the multi-objective approach cannot be avoided for robust
wind turbine control design.
1. For low wind speeds, to maximise the amount of the captured
wind power the blade pitch angle β must be held at a fixed
angle corresponding to maximum allowable conversion effi-
ciency curve at which β ¼0. In addition, the rotor speed must
vary in proportional to the wind speed variation so that λ is
4. The active sensor FTTC strategy
kept in the vicinity of its optimal value λopt (i.e. track the
optimal rotor rotational speed signal ωr opt ¼ ðλopt v=RÞ ). Speci-
This section focuses on the presentation of a combination of
fically, the generator subsystem represents the actuator of the
T–S fuzzy PPIO and TSDOFC for wind turbine sensor FTTC based on
aerodynamic subsystem in the low range of wind speed that
the scheme shown in Fig. 1. The advantageous features offered by
decelerates or just releases the aerodynamic subsystem rota-
the proposed strategy is that the T–S fuzzy PPIO and the TSDOFC
tion to adjust the variation of rotor speed, so that good tracking
are independently designed, and their performances are consid-
of the ωr opt ensured. It should be noted that exact tracking of
ered simultaneously, which is convenient for calculating the
the ωr opt leads to increase the load on the drive train shafts and
design parameters and can avoid the coupling problem generated
hence minimises the drive train life time. Moreover, exact
by the observer-based state feedback control (see Remark 3).
tracking will also produce a highly fluctuating output power
and produce a varying direction reference torque signals that
can lead to abnormal generator operation (Munteanu et al., Remark 3. In the fuzzy control design, each “local control” is
2008). designed from the corresponding local linear model of a T–S fuzzy
2. For high wind speeds, it is possible to dissipate some propor- model and the fuzzy controller design problem is to determine the
tion of the available wind power by changing the blade pitch local feedback gains within a parallel distributed compensation
angle to prevent the wind turbine operation from crossing over structure (Tanaka and Wang, 2001). Although the fuzzy controller
the rated power (i.e. regulate wind turbine operation at the is constructed using the local design structure, the feedback gains
rated power). However, to ensure good regulation performance, should be determined using global design conditions. It is reported
some control strategies use the generator torque control as a in Tanaka and Wang (2001) that for the fuzzy state estimate
supplementary control signal to overcome the limited rate of feedback control “the global design conditions are needed to
change of blade pitch actuator. guarantee the global stability and control performance”. Hence, the
fuzzy control designer does not have freedom to assign the local
The λopt is determined by relating the blade time tb and the system closed-loop poles anywhere in the stable complex plane.
turbulence time tw. tb is the time required by the blade to take the Therefore, the observer-based T–S state feedback control system
position of the previous one and tw is the time required to remove suffers a major drawback in that the observer dynamics may not
the disturbed wind component generated by the movement of the be assigned freely to satisfy closed-loop performance require-
blade. According to these two times, three operation scenarios can ments (i.e. the separation principle cannot be ensured even when
be recognised (Sami, 2012), these are the model uncertainty is not considered). On the other hand, the
sensor faults proposed in the benchmark model have an abrupt
a. t b 4 t w : This scenario corresponds to slow rotation in which change behaviour (Odgaard et al., 2013) for which the use of fast
some undisturbed wind passes the area swept by the rotor fault estimation observer is of great advantage. Hence, the combi-
without harvesting its power content. nation of T–S PPIO and TSDOFC is proposed in this strategy to
b. t b o t w : This scenario corresponds to fast rotation in which the overcome the limitation of T–S observer-based state feedback
blade passes through the disturbed wind component generated control.
6 M.S. Shaker, R.J. Patton / Engineering Applications of Artificial Intelligence 34 (2014) 1–12

4.1. The T–S fuzzy PPIO design observer gains to be designed, and ex is the state estimation error
defined as
Owing to the fact that obtaining accurate fault estimate signals
ex ¼ x  x^ ð16Þ
automatically implies FDD, the FDD and FTC literatures have shown a
constant increase in interest in developing observers that provide The estimator in Eq. (15) provides simultaneous estimation of
estimates of state and fault signals simultaneously. Within this frame- system state and fault signal.
work, proportional state estimator augmented with integral term (PIO)
has been developed as an extension to the well-known Luenberger Remark 4. The observability of Eq. (11) implies the observability
observer. Specifically, the PIO are observers in which an additional of the augmented system (15), this can be easily proved from the
term, proportional to the integral of the output estimation error, is observability condition:
2 3
added in order to provide estimates of constant or slowly varying fault sI  AðpÞ 0
signals. 6 7
rank4 As C sI þAs 5 ¼ n þl 8 s A ℂ
Since, in the fault estimation and compensation based FTC, a
0 I
post-fault system performance is highly affected by the estimation
accuracy of the fault signal and, on the other hand, the fault is an
unpredictable event in both its occurring time and its behaviour, Also, the rank condition is achieved since C Df ¼ As Df and As is
the fault estimation strategy must has due care for the time invertible, then rankðC Df Þ ¼ rankðAs Df Þ ¼ g.
varying nature of the fault.
The state estimation error dynamic then:
To enhance fault estimate capability of the PIO (i.e. provide accurate
estimation for time-varying fault scenarios), T–S fuzzy proportional e_ x ¼ ðAðpÞ  LðpÞCÞex þ Df ef þ EðpÞev ð17Þ
state estimators augmented with proportional and integral feedback
fault estimators (T–S fuzzy PPIO) is presented in this section. The ev is the difference between the effective wind speed and the
augmented proportional plus integral fault estimator provides degrees anemometer measured wind speed. Using Eqs. (15) and (17) the
of design freedom to shape the estimator tracking performance and fault estimation error dynamics are then as follows:
hence enhances the FTC loop performance. e_ f ¼ f_ s FðpÞCðAðpÞ  LðpÞC þ sÞex FðpÞC Df ef  FðpÞC EðpÞev ð18Þ
Consider the wind turbine system with sensor fault signal as
follows: The augmented estimator will then be of the following form:
)
x_ ¼ AðpÞx þ Bu þ EðpÞv e_~ a ðtÞ ¼ Aðp;
~ pÞe~ a þ N~ ðp; pÞz~ ð19Þ
ð11Þ
y ¼ Cx þ Df f s 2 3
AðpÞ  LðpÞC Df
The system matrices AðpÞ A ℛ nnn
ð ¼ ∑ri ¼ 1 hi ðpÞAi Þ,
BAℛ , EðpÞ A nnm ~
Aðp; pÞ ¼ 4 5
 FðpÞCðAðpÞ  LðpÞC þ IÞ  FðpÞC Df
ℛnnmv ð ¼ Σ i ¼ 1 hi ðpÞEi Þ, Df A ℛ and C A ℛ are known, r is the
r lng lnn

number of fuzzy rules, f s ¼ ½f sr f sg T A ℛg , and the term hi ðpÞ is " # " # " #
the weighting function satisfying Σ i ¼ 1 hi ðpÞ ¼ 1, and 1 Z hi ðpÞ Z 0;
r ex ev
e~ a ¼ ; z~ ¼ ; ~ ðp; pÞ ¼ EðpÞ
N
0
for all t. ef f_ s  FðpÞC EðpÞ I
Let ef A ℛg be the fault estimation error defined as
The objective is to compute the gains LðpÞ and FðpÞ such that the
ef ¼ f s  f^ s ð12Þ exogenous input z~ in Eq. (19) are attenuated below the desired
To avoid the direct multiplication of the sensor and/or noise by the level γ to ensure robust regulation performance in addition to
observer gain, an augmented system state with output filter states locate the observer poles within a specified disc region charac-
is constructed. The filtered output is given as follows: terised by its radius (α) and centre (β).

x_ s ¼  As xs þ As Cx þ As Df f s ð13Þ Remark 5. In the synthesis of control system, some desired


lnl performances should be considered in addition to stability. Since
where  As A R is a stable matrix. The augmented state system is
classical stability conditions do not deal with transient responses
given as
) of the closed-loop system, a satisfactory transient response can be
x_ ¼ AðpÞx þ Bu þ EðpÞv þ Df f s guaranteed by confining its poles in a prescribed region. For many
ð14Þ
y¼C x real problems, exact pole assignment may not be necessary and it
suffices to locate the closed-loop poles in a prescribed subregion in
" # " #   the complex plane. Within this framework Chilali and Gahinet
AðpÞ 0 xf B
AðpÞ ¼ ; x¼ ; B¼ (1996) proposed that a convex region in splane which represents
As C As xs 0
the desired closed-loop pole-placement constraints can be repre-
  " # sented as LMI.
EðpÞ 0
EðpÞ ¼ ; Df ¼ ; C ¼ ½0 Il  Consequently, in this paper, the transient performance of the
0 As Df
nonlinear system represented by T–S model is governed by
To deal with time-varying fault scenarios and perform fast fault applying the Chilali and Gahinet's LMI regions to each local T–S
estimation. A fuzzy fast adaptive fault estimator is designed as an fuzzy controller. Hence, for each local pair of AðpÞ and C matrices
extension to the work of Zhang et al. (2008) for linear systems. there will be observer gain LðpÞ (obtained offline during the design
Hence, the following fuzzy observer is proposed to simultaneously stage) that satisfies the feasibility of global stability as well as pole-
estimate the system states and sensor fault: placement LMIs.
9
_
x^ ¼ AðpÞx^ þ Bu þ Df f^ s þ EðpÞv~ þ LðpÞðCx  C xÞ
^ =
Theorem 1. The estimation error system eigenvalues are located in a
ð15Þ
_
f^ s ðtÞ ¼ FðpÞCðe_ x þex Þ ; disc region in the complex plane defined by (α and β) so that the
error dynamics are stable. Furthermore, the H1 performance is
where x^ A ℛn þ l is the estimation of the state vector x, v~ is anem- guaranteed with an attenuation level γ (provided that the signal
ometer measured wind speed, LðpÞ A ℛðn þ lÞl , and FðpÞ A ℛgl is the ðf_ s Þ is bounded), if there exist symmetric positive definite matrix P1,
M.S. Shaker, R.J. Patton / Engineering Applications of Artificial Intelligence 34 (2014) 1–12 7

matrices H i ; F i , and a scalar μ, α, and β satisfying the following LMI To conform to the format of P is structured as follows:
constraints:  
2 3 P1 0
 αP 1 0 ϕ1 ϕ2 P¼ 40 ð27Þ
0 I
6 0  αI ϕ3 ϕ4 7
6 7
6 7o0 ð20Þ
4 n n  αP 1 0 5 ~ ~
Substituting the corresponding values of P; Aðp; pÞ, Nðp; pÞ and
n n 0  αI using the variable change HðpÞ ¼ P 1 LðpÞ; and equality
2 3 T
Ψ 11 Ψ 12 Ψ 13 0 C Tp1 0 FðpÞC ¼ Df P 1 ð28Þ
6 7
6 n Ψ 22 Ψ 23 C Tp27
6 I 0 7 the LMI in Eq. (21) is thus obtained.
6 7
6 n n  γI 0 0 0 7
6 7 o0 ð21Þ To prove LMI (20) we first need the following lemma from
6 7
6 n n n  γI 0 0 7 Chilali and Gahinet (1996) and Mansouri et al.(2008).
6 7
6 n n n n  γI 0 7
4 5
n n n n n  γI Lemma 1. The matrix A is D-stable if and only if there exists a
symmetric matrix X such that
" #
LðpÞ ¼ P 1 1 HðpÞ; Ψ 11 ¼ P 1 AðpÞ þ ðP 1 AðpÞÞT  HðpÞC  ðHðpÞCÞT  αX βX þ XA
o 0; X 4 0 ð29Þ
T T T βX þ ðXAÞT  αX
Ψ 12 ¼ ðA ðpÞP 1 Df  C H ðpÞDf Þ; Ψ13 ¼ P 1 EðpÞ;
T T
Ψ22 ¼  2Df P 1 Df ; Ψ23 ¼  Df P 1 EðpÞ
Based on Lemma 1 one can obtain inequality (20) after
ϕ1 ¼ P 1 AðpÞ  HðpÞC þ βP 1 ; ϕ2 ¼ P 1 Df ; substituting
 
P1 0 ~
T T T T X¼P¼ 4 0; A ¼ Aðp; pÞ
ϕ3 ¼  ðA ðpÞP 1 Df  C H ðpÞDf þ Df P 1 ÞT 0 I

T
defined in (24).
ϕ4 ¼  D f P 1 D f þ β I This completes the proof of Theorem 1.

Proof. Let the performance output be defined as follows: 4.2. The TSDOFC design
" #
C p1 0
~e p ¼ C p e~ a ; C p ¼ The control objective here is to design a dynamic output
0 C p2 feedback controller capable of forcing the generator rotational
speed of the wind turbine to follow the optimal generator speed
signal in both faulty and fault-free cases.
C p1 A ℛknn and C p2 A ℛgng . The estimation performance objec-
An augmented system consisting of the system (11) and the
tive can be presented mathematically as follows: R
integral of the tracking error eti ¼ ðyr  SyÞ is defined as
jje~ p jj Z 1 Z 1 9
1
x_ ¼ AðpÞx þ Bu þEðpÞδ þ Ryr þ Din ef =
T
rγ ¼ e~ Ta C p C p e~ a dt  γ z~ z~ r 0
T
2
ð22Þ
jj z~ jj 2 γ 0 0 ð30Þ
y ¼ C x þ Df ef ;

Consider the following candidate Lyapunov function for the " #    


0 SC eti 0
augmented system: AðpÞ ¼ ; x¼ ; B¼
0 AðpÞ x B
υðe~ a Þ ¼ e~ Ta P e~ a ; where P 4 0 ð23Þ
" #
   
 SDf 0 I
Din ¼ ; EðpÞ ¼ ; R¼
To achieve the required performance (22) and the stability of 0 EðpÞ 0
the augmented system the following inequality should hold: " #
  0
I 0
1 C¼ ; Df ¼
υ_ ðe~ a Þ þ e~ Tp e~ p  γ z~ z~ o 0
T
ð24Þ 0 C Df
γ
where υ_ ðe~ a Þ is the derivative of candidate Lyapunov function in where S A ℛqnl is used to define which output variable is con-
terms of, it will become: sidered to track the reference signal. Since the system in (11) has
common input and output matrices (B and C), the dynamic output
υ_ ðe~ a Þ ¼ e~ Ta ðA~ ðp; pÞP þ P Aðp;
~
T
pÞÞe~ a þ e~ Ta P N~ ðp; pÞz~ þ z~ N~ ðp; pÞP e~ a
T T
feedback controller used to stabilise and perform the tracking
ð25Þ objective is of linear parameterisation form and defined as
)
x_ c ¼ Ac ðpÞxc þ Bc ðpÞðSr yr  yÞ
ð31Þ
By using (25), and the Schur Complement theorem, inequality u ¼ C c ðpÞxc þ Dc ðpÞðSr yr  yÞ
(24) implies that the following inequality holds:
where xc is the state and Ac ðpÞ A ℛðn þ qÞðn þ qÞ , Bc ðpÞ A ℛðn þ qÞðl þ qÞ ,
C c ðpÞ A ℛmðn þ qÞ , Dc A ℛmðl þ qÞ and Sr A ℛðl þ qÞq is introduced to
2 3
match the dimensions of yr and y.
~
A~ ðp; pÞP þ P Aðp;
T
pÞ ~
P Nðp; pÞ C p1
T
0
6 7
6 T 7 Remark 6. In general, the TSDOFC can have cubic, quadratic, or
6n
6  γI 0 C p2 7
7o0 ð26Þ
6n linear parameterisation. For a given T–S model, the choice of
4 n  γI 0 7 5
particular TSDOFC parameterisation is influenced by the structure
n n n  γI
of the T–S subsystems. While cubic parameterisation is required
8 M.S. Shaker, R.J. Patton / Engineering Applications of Artificial Intelligence 34 (2014) 1–12

for parameter dependent A; B; and C matrices T–S fuzzy model, ^


Ac ðpÞ ¼ N  1 ðAðpÞ b
 YðAðpÞ  BDðpÞCÞX  YBC c ðpÞM T þ NBc ðpÞCXÞM  T
the linear parameterisation fuzzy controller is suitable for com-
mon input and common output matrices. where M and N satisfy MNT ¼ I  XY.

Aggregation of (30) and (31) gives the following system: Proof. Let the output performance yp be defined as
)
x_ a ¼ Aa ðpÞxa þ Ea ðpÞd yp ¼ C p y
ð32Þ
y ¼ C a xa þ Da d
where C p is selected according to design requirements. The
2 3 objective of the controller design performance can presented
" # δ
AðpÞ BDc ðpÞC BC c ðpÞ 6 7 mathematically as
Aa ðpÞ ¼ ; d ¼ 4 ef 5
Bc ðpÞC Ac ðpÞ jjyp jj Z 1 Z 1
yr 1 T
rγ ¼ yT C p C p ydt  γ
T
2
d d r0 ð35Þ
jjdjj2 γ 0 0
" #
x
xa ¼ ; C a ¼ ½C 0; Da ¼ ½0 Df 0
xc Consider the following candidate Lyapunov function for the
augmented system (32) :
" #
EðpÞ Din  BDc ðpÞDf R þ BDc ðpÞSr
Ea ðpÞ ¼ υðxa Þ ¼ xTa P c xa ; where P c 40
0  Bc ðpÞDf Bc ðpÞSr

To achieve the required performance (35) and stability of the


Theorem 2. The eigenvalues of the closed-loop system (32) are augmented system (32) the following inequality should hold:
located in the disc region of the negative complex plane characterised
by radius (α), centre (β), and the closed-loop is stable and tracks the 1
υ_ ðxa Þ þ yTp yp  γ dT d o 0 ð36Þ
reference signal with guaranteed H1 performance and with an γ
attenuation level γ (provided that the signals in d are bounded), if where υ_ ðxa Þ is the derivative of candidate Lyapunov function. In
there exist symmetric positive definite matrices X; Y and matrices terms of Eq. (32), this becomes:
Ac ðpÞ, Bc ðpÞ, C c ðpÞ, and Dc ðpÞ satisfying the following LMI constraints:
2 3 υ_ ðxa Þ ¼ xTa ðATa ðpÞP c þP c Aa ðpÞÞxa þ xTa P c Ea ðpÞd þdT ETa ðpÞP c xa ð37Þ
 αX I ϕ1c ϕ2c
6 I  αY ϕ3c ϕ4c 7
6 7
6 7o0 ð33Þ
4 n n  αX 0 5 By using Eq. (37), and the Schur Complement theorem, inequal-
n n 0  αY ity (36) implies that the following inequality must hold:
2 3
2 3 T
AT ðpÞP c þ P c Aa ðpÞ P c Ea ðpÞ C p
Ψ 11c Ψ 12c EðpÞ Ψ 13c Ψ 14c XC Tp1 6 a 7
6 7 6n  γI 0 7
6 n 4 5 o0 ð38Þ
6 Ψ 22c YEðpÞ Ψ 23c Ψ 24c C Tp1 7
7
6 7 n n  γI
6 n n  γI 0 0 0 7
6 7o0 ð34Þ
6 7
6 n n n  γI 0 0 7
6 7  
6 n n n n  γI 0 7 1 Y N
4 5 Assume P c and P c is structured as P c ¼ T ,
n n n n n  γI N W
    
1 X M 1 X M Y N
where Pc ¼ since P c P c ¼ I (i.e. ¼
MT Z MT Z NT W
" #      
^
Ψ11c ¼ AðpÞX þðAðpÞXÞT þ BCðpÞ ^
þ ðBCðpÞÞT
XY þ MN T
XN þ MW I 0 X I
¼ )We have P c ¼ )
M T Y þ ZN T M T N þ ZW 0 I MT 0
T    
Ψ 12c ¼ A^ ðpÞ þAðpÞ BDðpÞC
b X I I Y
Pc ¼ .
MT 0 0 NT
b
Ψ 13c ¼ Din  BðpÞDðpÞD f    
X I I Y
Define Π 1 ¼ ; Π2 ¼
b
Ψ 22c ¼ YAðpÞ þðYAðpÞÞT þ BðpÞC b
þ ðBðpÞCÞT M T
0 0 N T

Pre- and post-multiply inequality (38) by diagonals ½Π 1


T
I I
b
Ψ 23c ¼ YDin þ BðpÞD and its transpose and by using the variable change:
f

^
AðpÞ b
¼ YðAðpÞ  BDðpÞCÞX þ YBC c ðpÞM T  NBc ðpÞCX þ NAc ðpÞM T
b
Ψ 14c ¼ R þ BDðpÞS b
r ; Ψ 24c ¼ YR  BðpÞSr

b ¼  NBc ðpÞ  YBDc ðpÞÞ


BpÞ
^
ϕ1c ¼ AðpÞX þ BCðpÞ b
þ βX; ϕ2c ¼ AðpÞ  BDðpÞC þ βI

^ b ^
CðpÞ ¼ C c ðpÞM T  Dc ðpÞCX
ϕ3c ¼ AðpÞ þ βI; ϕ4c ¼ YAðpÞ þ BðpÞCðpÞ þ βY
The controller gains are thus calculated as follows: b
DðpÞ ¼ Dc ðpÞ
b
Dc ðpÞ ¼ DðpÞ
Inequality (34) can thus be easily obtained.
By using Eq. (29) of Lemma 1 and the variable change
^
C c ðpÞ ¼ ðCðpÞ þ Dc ðpÞCXÞM  T A ¼ Aa ðpÞ; X ¼ P c and then by pre- and post-multiplying by diag-
onal ½Π 1 Π 1  and its transpose respectively, the inequality (33)
T T
b  YBDc ðpÞ
Bc ðpÞ ¼ N  1 ð  BðpÞ can be easily obtained.
M.S. Shaker, R.J. Patton / Engineering Applications of Artificial Intelligence 34 (2014) 1–12 9

Remark 7. as the severity of the sensor fault increases the tracking


performance then deeply degrades.
 Equality (28) can be relaxed using the following optimisation  Stuck measurement sensor fault: ðωr measured ¼ constantÞ .This fault
problem (Zhang et al., 2008): scenario represents the worst sensor fault case since the controller
Minimise μ receives measurements that are completely independent of the
" T
# system states. In fact, this fault leads the controller to generate
μI Df P 1  FðpÞC control signal that forces the closed-loop system to operate away
40 ð39Þ
n μI from nominal operating region boundary since whatever control
signal is delivered, the measurements are fixed at C i xðtÞ ¼ constant
and hence ðet ¼ ωr opt  ωr measured a 0Þ.
Matrices M and N T can be calculated based on equality  Without loss of generality, the output matrix parametric fault
MN T ¼ I  XY using any matrix decomposition techniques such presented in wind turbine benchmark model can be repre-
as qr or svd. sented as an additive fault in which the fault signal depends on
 Practically, the measured wind speed signal does not exactly the measured state, as illustrated below:
represent the effective wind speed signal. However, the pro-
posed strategy makes use of measured wind speed in order to
Suppose the fault free output is given as
cope with the tradeoff between design complexity and wind
 " #
speed signal accuracy. However, the fault estimator is designed 1 0 x1
y ¼ Cx ¼
to be robust against the expected error between the measured 0 1 x2
and the effective rotor wind speed.
 Due to the large wind turbine inertia, in the practical situation
the tracking follows the trends of the speed variation and not Then the parametric sensor fault can be modelled as follows:
the detailed variations.  " #  
1 0 x1 0
yf ¼ C f x ¼ ¼ Cx þ ð  0:1n x2 Þ ð41Þ
0 0:9 x2 1

5. Simulation results
Three fault scenarios are considered in this section, these are
The FTTC system design and simulation results are based on the the following.
wind turbine benchmark system proposed by kk-electronic
(Odgaard et al., 2009, 2013). As illustrated in Section 2, the T–S 5.1. Generator rotational speed scaling fault ðf sg Þ
fuzzy model describing the low wind speed dynamics of the wind
turbine is derived depending on two scheduling variables, namely The two fault scenarios are 0.9 and 1.1 scale measurements (output
the measured rotor angular velocity ωr and the wind speed v. The matrix parametric changes) of generator rotational speed sensor. As
nonlinear system behaviour is approximated using four local stated in Remark 8, parameter changes in the output matrix C can be
linear systems derived to represent the system dynamics at four considered as a special case of additive faults in which the fault signals
operating points. (f sg ) is a scaled version of the measured state. Fig. 6 shows the two
As stated in Sections 3.3, the controller optimises the power fault scenarios and the effectiveness of the proposed strategy to
captured by controlling the rotor rotational speed by varying the compensate the bias from the scaled measurements.
reference generator torque Tgr so that the wind turbine rotor speed The ability of the proposed T–S fuzzy PPIO to accurately
ωr follows the optimal rotor speed given by estimate abruptly changing fault is clearly shown in Fig. 7(a and b).
The T–S PPIO can provide information about the fault severity
λopt v
ωr opt ¼ ð40Þ via the fault estimation signal. This is achieved through taking the
R
ratio between the measured generator speed ωg and the estimated
where ωr opt and λopt are the optimal rotor speed and the optimal tip signal ω ^ g . Hence, if there are no faults the ratio should be
speed ratio. In fact, designing a controller for the power optimisation 1 otherwise, any deviation from unity indicates the occurrence
problem must achieve the design constraints listed in Remark 2. One of the fault and the magnitude of the deviation represents the fault
of these constraints is to have the capability to tolerate the effects of severity. Fig. 8 shows the fault evaluation signal for both 1:1 ωg
faults that affect different system components. and 0.9ωg fault scenarios.
The following faults are considered and the proposed control Clearly, the two fault scenarios (1.1 or 0.9 scale fault) affect the
strategies need to tolerate the fault effects so that good tracking wind turbine closed-loop performance and hence the wind power
performance to ωr opt can be maintained. conversion efficiency. However, the expected effect of this fault is

Remark 8. The expected effects of different fault scenarios pro- ωg in different cases
100
posed in the considered benchmark is investigated as follows:
90 Bias gen. sensor fault

 Scaling measurement sensor fault: This fault scenario represents 80


ωg (rad/s)

a loss of effectiveness of the sensor so that the faulty measure- 70

ment becomes ωr measured ¼ α ωr actual . Since the tracking pro- 60

blem becomes a regulation problem in terms of the tracking 50


ωg with fault compensated
error ðet Þ. Suppose, following a transient time, that the tracking 40
ωg with 1.1 fault
error ðet Þ becomes ðet ¼ ωr opt  ωr measured ¼ 0Þ, this implies 30
ωg with 0.9 fault
that: ωr actual ¼ ωr measured =α. Hence, from a control stand point, 20
0 100 200 300 400 500 600
for the case of a sensor fault the controller starts to produce a
Time(s)
control signal that minimises the difference between the faulty
measurement and the desired optimal rotor speed. Moreover, Fig. 6. Effectiveness of compensation strategy.
10 M.S. Shaker, R.J. Patton / Engineering Applications of Artificial Intelligence 34 (2014) 1–12

Estimation of fsg Effect of 1.1 fsg


10 1
0.95 ω r without fsg
8
ω r with 1.1 fsg
6 0.9
fsg (rad/s)

Fault duration

ωr (rad/s)
0.85
4
0.8
2
0.75
0
0.7
-2 0.65
0 100 200 300 400 500 600
Time(s) 0 100 200 300 400 500 600
Time(s)
The estimation of fsg
2 Fig. 9. The effect of generator speed scale sensor fault.

-2 ω g in different cases
fsg (rad/s)

120
-4 1.5 ωg sensor fault

-6 100

ω g (rad/sec)
-8
80
-10
0 100 200 300 400 500 600
60
Time(s) Nominal ωg

Fig. 7. Generator rotational speed scale sensor faults estimation (a and b). 40 ωg with 1.5 fault
ωg with fault compensation
20
0 100 200 300 400 500 600
fsg evaluation
Time (sec)

1.15 The estimation of fsg


40
1.1
Fault

1.05
+0.1 30
fsg (rad/sec)

1
20
0.95
10
0.9
0 100 200 300 400 500 600
0
Time(s)
-10
fsg evaluation 0 100 200 300 400 500 600
1.1 Time (sec)
1.05 Fig. 10. Effectiveness of compensation strategy with 50% scale fault. (a) ωg in
different cases and (b) fault estimation.
1
Fault

-0.1
0.95

0.9 always receives a fault-free measurement. Hence, the ability of the


proposed FTTC to tolerate bigger fault magnitudes is governed by the
0.85
ability of T–S fuzzy PPIO to provide a fault estimation signal. This in turn
0.8 depends on whether the post-fault pairs ðAðpÞ; and CÞ satisfy the
0 100 200 300 400 500 600
observability condition (or at least the detectability condition). Fig. 10a
Time(s)
shows the generator rotational speed with 1:5 ωg scale fault. Although
Fig. 8. Deviation of 1.1 (a) and 0.9 (b) sensor measurements from unity. this fault has bigger magnitude than 1:1 ωg , the T–S fuzzy PPIO can
provide an accurate fault estimate (see Fig. 10b) which in turn used to
probably less than the effect of rotor speed sensor fault since the compensate (hide) the effect of 1:5 ωg sensor fault from the input of
generator speed signal is part of the feedback signals and not TSDOFC. This is clearly shown in Fig. 10a in which the measured ωg after
compared directly with the reference optimal speed (i.e. not the fault compensation exactly match the nominal (fault free) ωg.
objective signal). Fig. 9 shows the tracking of the objective variable ωr
with(out) 1:1 ωg scaling fault.
Although the ωg sensor fault has minor effect on the overall 5.2. Rotor rotational speed stuck sensor fault ðf sr Þ
closed-loop system performance, the proposed strategy compen- (ωr opt o ωr measured )
sates the effect of this fault scenario so that the TSDOFC always
receives fault free ωg measurement. The second fault is represented by the fixed sensor output of
the rotor speed sensor at1:4 rad s  1 . This fault scenario and the
Remark 9. As presented in Section 2, the proposed FTTC is based on effectiveness of the estimation and compensation strategy to
the estimation and compensation approach to FTC. Within this frame- maintain the required system performance in the presence of this
work the estimator hides the effect of sensor faults so that the controller fault are shown in Fig. 11.
M.S. Shaker, R.J. Patton / Engineering Applications of Artificial Intelligence 34 (2014) 1–12 11

ωr in different cases ωr in different cases


1.5 2

1.5

1
1

ωr (rad/s)
ω r (rad/s)

0.5
Turbine cut-off limit
0
0.5 -0.5
ω r with 1.4 constant fsr ω r with 1.4 constant fsr
-1
ω r with fsr compensated ω r without fsr compensation
0 -1.5
0 100 200 300 400 500 600 0 100 200 300 400 500 600

Time(s) Time(s)

The estimation of f sr Tg in different cases


0.8 8000
Tgr without fsr compensation
0.6 7000
Tgr with fsr compensation
0.4 6000
fsr (rad/s)

0.2 5000

Tgr (Nm)
0 4000
-0.2 3000
Fault estimation
-0.4 2000

-0.6 1000

-0.8 0
0 100 200 300 400 500 600 0 100 200 300 400 500 600

Time(s) Time(s)

Fig. 11. The effectiveness of the proposed strategy to tolerate stuck rotor speed Fig. 12. Further investigation of stuck fault effect and the effectiveness of the
sensor fault (a) and fault estimation (b). proposed strategy: (a) rotor speed and (b) generator torque.

The effect of this fault scenario varies according to the fixed ω r in different cases
1.5
measurement of rotor speed and ωr opt which in turn depends on the
wind speed. In this case ωr opt is lower than the fixed rotor speed
measurement and hence the controller starts to force the system to 1
ωr (rad/s)

slow-down the rotor speed (Fig. 12a). Hence, as long as the optimal
speed remains below the measured value, the controller keeps
increasing the reference generator torque Tgr (Fig. 12b) which may 0.5
Measured ω r
lead to the rotation speed ωr reaching its cut-off value, i.e. the
ω r without fsr comp.
turbine is shut down due to a rotor rotation speed sensor fault. ω r with fsr comp.
A further investigation of the effect of stuck sensor fault scenario is 0
0 100 200 300 400 500 600
shown in Fig. 12. It is clearly shown that this fault scenario can lead
Time(s)
to wind turbine shut-down by increasing the breaking action which
in turn increases the drive train torsional load. Tg in different cases
3000

5.3. Rotor rotational speed stuck sensor fault (ωr opt 4 ωr measured ) 2500

2000
Tgr (rad/s)

In this rotor speed stuck sensor fault scenario ωr opt is higher


than the stuck measurement (at 0:5 rad s  1 ) of the rotor speed. 1500

The effectiveness of the proposed strategy to tolerate this fault 1000


scenario is shown in Fig. 13(a and b). In this case, the controller
500 Tgr with fsr compensation
will generate a reference generator torque signal (Tgr) that simply Tgr without fsr compensation
release the turbine to rotate according to the available wind speed 0
0 100 200 300 400 500 600
without control (Fig. 13b).
On the other hand, Fig. 14 shows the effect of stuck fault Time(s)
scenarios from power capture stand point. For 1:4 rad s  1 stuck Fig. 13. The effectiveness of the proposed strategy to tolerate stuck rotor speed
sensor fault the controller will produce Tgr that forces the wind sensor fault: (a) rotor speed and (b) generator torque.
turbine to slow down the actual rotor rotational speed and hence
minimises the power capture (Fig. 14 dotted line). On the other
hand, 0:5 rad s  1 stuck fault will make the controller produces Tgr proposed architecture, a detailed design approach is presented for
that simply release the wind turbine to rotate according to the an OWT. A dynamic output feedback control scheme is used that
available wind speed. This in turn will forces the system to operate has time-varying reference tracking capability and the fault
away from its optimal tip speed ratio (Fig. 14 dashed-dotted line). estimator is designed using a proposed T–S extension of a well
known PPI observer scheme, the T–S PPIO. The controller and fault
estimator individually satisfy an appropriate L2 norm robustness
6. Conclusion condition guaranteeing minimum tracking error and robust fault
estimation. Results are presented illustrating the robustness of the
The paper develops a new architecture for active sensor FTTC proposed FTTC design using the Safeprocess wind turbine bench-
for OWT based on fault estimation and compensation. Using this mark system.
12 M.S. Shaker, R.J. Patton / Engineering Applications of Artificial Intelligence 34 (2014) 1–12

x 105 Pcap in different fsr scenarios Badihi, H., Zhang, Y., Hong, H., 2013. Fuzzy gain-scheduled active fault-tolerant
20 control of a wind turbine. J. Frankl. Inst. http://dx.doi.org/10.1016/j.jfranklin.
2013.05.007, in press.
15 Bianchi, D.F., De Battista, H., Mantz, J.R., 2007. Wind Turbine Control Systems:
Principles, Modelling and Gain Scheduling Design. Springer-Verlag, London.
Pcap (W)

10 Chilali, M., Gahinet, P., 1996. H1 design with pole placement constraints: an LMI
approach. IEEE Trans. Autom. Control 41 (3), 358–367.
5 Djurovic, S., Crabtree, C.J., Tavner, P.J., Smith, A.C., 2012. Condition monitoring of
wind turbine induction generators with rotor electrical asymmetry. Renew.
Pcap with fsr compensation
Power Gener.: IET 6 (4), 207–216.
0 Pcap with 0.5 const fsr Hameed, Z., Hong, Y.S., Cho, Y.M., Ahn, S.H., Song, C.K., 2009. Condition monitoring
Pcap with 1.4 const fsr and fault detection of wind turbines and related algorithms: a review. Renew.
-5 Sustain. Energy Rev. 13 (1), 1–39.
0 100 200 300 400 500 600
Jain, T., Yame, J.J., Sauter, D., 2013. A novel approach to real-time fault accommoda-
Time (s) tion in NREL’s 5-MW wind turbine systems. IEEE Trans. Sustain. Energy 4 (4),
1082–1090.
Fig. 14. The effect of stuck fault scenarios on the power captured. Johnson, K.E., Fleming, P.A., 2011. Development, implementation, and testing of
fault detection strategies on the National Wind Technology Center’s controls
advanced research turbines. Mechatronics 21 (4), 728–736.
Kamal, E., Aitouche, A., Ghorbani, R., Bayart, M., 2012. Robust fuzzy fault-tolerant
In summary, wind turbines are nonlinear systems that are control of wind energy conversion systems subject to sensor faults. IEEE Trans.
driven by a stochastic wind force signal. Sustainability of such Sustain. Energy 3 (2), 231–241.
systems is highly affected by the chosen control strategy. Several Mansouri, B., Manamanni, N., Guelton, K., Djemai, M., 2008. Robust pole placement
controller design in LMI region for uncertain and disturbed switched systems.
design constraints must be taken into account in the design of the
Nonlinear Anal.: Hybrid Syst. 2 (4), 1136–1143.
wind turbine power maximisation controller, these are Marden, J.R., Ruben, S.D., Pao, L.Y., 2013. A model-free approach to wind farm
control using game theoretic methods. IEEE Trans. Control Syst. Technol. 21 (4),
a. Wind turbines are characterised by their non-linear aerody- 1207–1214.
Munteanu, I., Bratcu, A., Cutululis, N.-A., Ceanga, E., 2008. Optimal Control of Wind
namics and have a stochastic and uncontrollable driving force Energy Systems: Towards a Global Approach. Springer-Verlag, London.
as input in the form of wind speed. This limits the ability of Niemann, H., Stoustrup, J., 2005. Passive fault tolerant control of a double inverted
linear control strategies to maintain acceptable performance pendulum—a case study. Control Eng. Pract. 13 (8), 1047–1059.
Odgaard, P.F., Stoustrup, J., Kinnaert, M., 2009. Fault tolerant control of wind
over a wide range of wind speed. turbines: a benchmark model. In: Proceedings of the 7th IFAC Symposium on
b. Due to the common input common output matrices of wind Fault Detection,Supervision and Safety of Technical Processes Safeprocess 2009,
turbine model the conservatism of T–S fuzzy estimation and Barcelona. June 30–July 3, pp. 155–160.
Odgaard, P.F., Stoustrup, J., Kinnaert, M., 2013. Fault-tolerant control of wind
control is highly reduced.
turbines: a benchmark model. IEEE Trans. Control Syst. Technol. 21 (4),
c. Owing to the direct effect of wind turbine components faults on 1168–1182.
the wind power conversion efficiency, the designed control Sami, M. 2012. Active Fault-Tolerant Control of Nonlinear Systems with Wind
strategy must be capable of tolerating different expected fault Turbine Application (Ph.D. thesis). The University of Hull.
Sami, M., Patton, R.J., 2012a. Fault tolerant adaptive sliding mode controller for
effects. wind turbine power maximisation. In: Proceedings of the 7th IFAC Symposium
d. Exact tracking leads to increased loading on the two drive train on Robust Control Design,Aalborg Congress & Culture Centre, Denmark. 20–22
shafts and hence can shorten the drive train life time. This also June, pp. 499–504.
Sami, M., Patton, R.J., 2012b. Wind turbine power maximisation based on adaptive
produces a highly fluctuating output power, and may even sensor fault tolerant sliding mode control. In: Proceedings of the 20th
produce a varying direction reference torque signal that can Mediterranean Conference on Control & AutomationBarcelona. 3–6 July,
lead to abnormal generator operation. It is thus very clear that pp. 1183–1188.
Simani, S., Castaldi, P., 2014. Active actuator fault-tolerant control of a wind turbine
the multi-objective approach cannot be avoided for robust
benchmark model. Int. J. of Robust Nonlinear Control. 24 (8–9), 1283–1303.
wind turbine control design. Simani, S., Castaldi, P., 2013b. Data-driven and adaptive control applications to a
e. Due to the stochastic nature of the operating environment and wind turbine benchmark model. Control Eng. Pract. 21 (12), 1678–1693.
Sloth, C., Esbensen, T., Stoustrup, J., 2011. Robust and fault-tolerant linear
missing or inaccurate wind speed measurements, the wind
parameter-varying control of wind turbines. Mechatronics 21 (4), 645–659.
turbine fault estimation and compensation problem are chal- Stewart, G., Lackner, M., 2013. Offshore wind turbine load reduction employing
lenged by a requirement for a robust against the difference optimal passive tuned mass damping systems. IEEE Trans. Control Syst.
between the measured and the actual wind speed. Technol. 21 (4), 1090–1104.
Tanaka, K., Wang, H.O., 2001. Fuzzy Control Systems Design and Analysis: A Linear
Matrix Inequality Approach. John Wiley & Sons, New York, USA.
Verbruggen, T.W., 2003. Wind Turbine Operation and Maintenance Based on
References Condition Monitoring. Technical Report ECN-C-03-047. Energy Research Center,
the Netherlands.
Wei, X., Liu, L. 2010. Fault estimation of large scale wind turbine systems. In:
Amirat, Y., Benbouzid, M.E.H., Al-Ahmar, E., Bensaker, B., Turri, S., 2009. A brief Proceedings of the 29th Chinese Control Conference. 29–31 July, pp. 4869–4874.
status on condition monitoring and fault diagnosis in wind energy conversion Zhang, K., Jiang, B., Cocquempot, V., 2008. Adaptive observer-based fast fault
systems. Renew. Sustain. Energy Rev. 13 (9), 2629–2636. estimation. Int. J. Control Autom. Syst. 6 (3), 320–326.

You might also like