Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Information Sciences 582 (2022) 73–88

Contents lists available at ScienceDirect

Information Sciences
journal homepage: www.elsevier.com/locate/ins

Fault-tolerant load frequency control for DFIG-based


interconnected wind power systems
Subramanian Kuppusamy a, Young Hoon Joo a,⇑, Han Sol Kim b
a
School of IT Information and Control Engineering, Kunsan National University, 588 Daehak-ro, Gunsan-si, Jeonbuk 54150, Republic of Korea
b
Department of Control and Automation Engineering, Korea Maritime and Ocean University, Busan, Republic of Korea

a r t i c l e i n f o a b s t r a c t

Article history: This paper investigates the fault-tolerant load frequency control (LFC) design of intercon-
Received 1 April 2021 nected wind power systems. For this, a doubly-fed induction generator (DFIG)-based wind
Received in revised form 2 September 2021 farm is integrated into the interconnected power systems. Distinct from the existing works,
Accepted 4 September 2021
the communication time-delay is taken into the area control error (ACE) signals of inter-
Available online 08 September 2021
connected wind power systems. Also, compared with the traditional LFC scheme, a fault-
tolerant H1 LFC scheme is presented with known and unknown actuator fault against load
Keywords:
disturbances and wind speed fluctuations. Then, the asymptotic stability conditions are
Fault-tolerant control
Interconnected power systems
derived in the linear matrix inequalities (LMIs) form by utilizing the Lyapunov–
Load frequency control Krasovskii functional (LKF) method and Wirtinger-based inequality technique under the
Linear matrix inequality designed LFC scheme with H1 performance. Finally, numerical examples are given to val-
idate the applicability and superiority of the designed LFC scheme.
Ó 2021 Elsevier Inc. All rights reserved.

1. Introduction

Power systems are interconnected systems, and it has interconnection among generators, transformers, transmission
lines, loads, and much other protective equipments [1–4]. When increasing the power demands, renewable energy power
sources, such as wind, geothermal, photovoltaic, etc., are integrated with the power systems. Compared to other renewable
energies, wind power has gained considerable interest as a zero-emission sustainable renewable energy, and it is easily
extracted from the wind through wind turbines to produce electrical power [5–7]. For example, the power system including
a wind farm system is proposed in [8] using an energy capacitor system to minimize the frequency fluctuation.
On the other hand, the impact of wind turbine generators on power system dynamic behavior will differ from conven-
tional synchronous generators. Thus, it is significant to analyze the dynamic behavior of wind turbine models in power sys-
tems [9–11]. Recently, in the grid-connected and standalone wind power systems, DFIGs have been widely utilized since it
has effective four-quadrant active and reactive power capability, flexibility for variable-speed wind turbines [12–14]. In
addition, the fast power decoupling control capabilities of DFIGs can significantly inhibit voltage changes under wind speed
variations. For example, a coordinated four-loop switching controller of DFIG has been designed to improve the transient
stability of wind power penetrated power systems in [15]. Moreover, based on the input-to-state stability approach, the fre-
quency response has been analyzed in [16] by using DFIG’s output power integrated into the dynamic model of the power
grid.

⇑ Corresponding author.
E-mail address: yhjoo@kunsan.ac.kr (Y.H. Joo).

https://doi.org/10.1016/j.ins.2021.09.003
0020-0255/Ó 2021 Elsevier Inc. All rights reserved.
S. Kuppusamy, Young Hoon Joo and Han Sol Kim Information Sciences 582 (2022) 73–88

With the integration of wind power, the frequency preserving to the planned limit is essential for active power balance in
an area. Apart from that, the power systems may become unstable. At the same time, in the design and operation of inter-
connected wind power systems, the LFC has received great attention from the researchers, since LFC is used to maintain the
system frequency at a rated value and ensure power quality [17,19,18,20,21]. From this, the stability conditions for multi-
area power systems have been derived in [17] via the LFC scheme. The LFC has been designed in [18] for hybrid power sys-
tems through the disturbance observer approach. Recently, the H1 LFC problem has been studied for the multi-area power
systems under the hybrid cyber-attacks in [19]. Hence, the design of LFC is a significant factor to guarantee the safety and
stability of interconnected wind power systems. To mention a few, the LFC has been designed in [22] for multi-area time-
delay power system with wind power integration via the sliding mode method. Moreover, the authors have addressed
the LFC design problem for interconnected power systems with wind turbines in [23].
It is worth noting that the signal transmission among the various control areas in interconnected power systems induced
the ignorable time-delay and it may cause poor dynamic performance or instability of LFC systems. For instance, the H1 LFC
design for interconnected power systems with communication delay has been established in [24]. By including the commu-
nication delays, the LFC design problem for an interconnected power system which incorporates the effect of using an open
communication network for the ACE signals has been analyzed in [25]. The time-delay has occurred in the uncertain inter-
connected power systems because of the information transfer over the communication networks [26], and the LFC has been
designed to achieve the stability of the overall closed-loop system.
Fault-tolerant control (FTC) is a powerful tool to compensate for actuator faults and to ensure the reliability and safety of
dynamic systems, see, [27–34]. The asymptotic stability problem for the uncertain networked control system has been inves-
tigated under the FTC with known and unknown actuator faults in [27]. In [30], the FTC has been designed for networked
control systems subject to actuator fault under event-triggering sampling strategy. Recently, the FTC plays a significant role
in controlling wind power systems. For example, the stabilization problem of the DFIG-based wind turbine system has been
addressed in [31] via FTC with stochastic actuator faults. The LFC design for uncertain multi-area power systems with con-
stant time-delay has been studied in [35]. However, these methods only pay attention to interconnected power systems or
DFIG-based wind turbine systems separately and are not focused on the DFIG-based wind farm integrated into the power
systems, which is the main motivation of this work.
From the aforementioned works, we would like to mention that the fault-tolerant LFC design of interconnected wind
power systems has not been thoroughly studied. To fill these gaps in existing studies, we design the fault-tolerant LFC for
the first time of interconnected wind power systems, including the communication delays, load disturbances, and wind
speed fluctuations under known and unknown actuator fault cases. In the following, the delayed feedback and state feedback
H1 LFC are designed which ensure the asymptotic stability for interconnected delayed power systems. Finally, in the numer-
ical examples, the designed control scheme is validated through the interconnected wind power systems, and the effective-
ness of proposed method is demonstrated via two-area power systems with communication delays.
The overall structure of the paper is given as follows. Section 2 formulates the interconnected wind power systems. The
fault-tolerant LFC is designed for interconnected wind power systems in Section 3. The H1 LFC design for two-area power
systems is given in Section 4. Section 5 provides numerical examples. Finally, Section 6 presents the conclusion.
Notations: Rmn represents the set of all m  n real matrices. SymfBg ¼ B þ BT . The matrix B > 0 ð< 0Þ is a positive (neg-
ative) definite. The n  n identity matrix and zero matrix are represented by Inn and 0nn , respectively. diagf  g denotes a
block diagonal matrix. eTi ¼ ½0nði1Þn Inn 0nð10iÞn  ði ¼ 1; 2; . . . ; 10Þ.

2. Problem formulation

2.1. Mathematical model of DFIG

The mathematical model of the voltage equations for DFIG can be described as
8
dWds
>
> V ds ¼ Rs ids  Wqs þ x1s
>
> dt
>
>
< V qs ¼ Rs iqs  Wds þ 1 dWqs
xs dt
ð1Þ
>
> V dr ¼ Rr idr  sWqs þ x1s dWdr
>
> dt
>
>
: V ¼ R i þ sW þ 1 dWqr
qr r qr dr xs dt

where V ds ; V qs and V dr ; V qr are the d  q axis components for stator, and injected rotor voltage; ids ; iqs and idr ; iqr are the d  q
axis components for stator, and rotor current; s represents the rotor slip; the electrical speed of air–gap flux is denoted as
xs ; Rs and Rr indicate the resistance of stator and rotor, respectively; W is the flux linkage. The equation of flux linkage
can be written as

74
S. Kuppusamy, Young Hoon Joo and Han Sol Kim Information Sciences 582 (2022) 73–88

8
>
>
Wds ¼ ðLs þ Lm Þids þ Lm idr
>
< W ¼ ðL þ L Þi þ L i
qs s m qs m qr
ð2Þ
> Wdr ¼ ðLr þ Lm Þidr þ Lm ids
>
>
:
Wqr ¼ ðLr þ Lm Þiqr þ Lm iqs
where Ls ; Lr , and Lm denote the stator, rotor leakage inductance, and magnetizing inductance of the generator. The equation
of the electromagnetic torque T e can be described as
T e ¼ Wds iqs  Wqs ids ¼ Wdr iqr  Wqr idr : ð3Þ
From (1) and (2), we get the following
8
< i_dr ¼ xs ðV dr  Rr idr Þ þ sxs ðiqr þ Lm iqs Þ  Lm i_ds
Lr þLm Lr þLm Lr þLm
ð4Þ
: i_qr ¼ xs ðV qr  Rr iqr Þ  sxs ðidr þ Lm ids Þ  Lm i_qs
Lr þLm Lr þLm Lr þLm

Choose Wds ¼ 1p:u, and Wqs ¼ 0, then the following equation is obtained from (2)
Lm
iqs ¼  iqr :
Ls þ Lm
2
Set X 3 ¼ LsLþL
m
m
; X 2 ¼ R1r ; L1 ¼ Lr þ Lm þ LsLþL
m
m
and T 1 ¼ xLs1Rs . Based on the Laplace transform in (4), and applying simple math-
ematical calculation, one can get
8
< i_dr ¼  1 iqr þ X 2 V qr
T1 T1
:x
_ s ¼  2M
X3
iqr þ 2M
1
Tm
t t

Based on the above, the electromagnetic torque (3) can be rewritten as


T e ¼ iqs ¼ X 3 iqr :
The output wind power is represented as follows:
P e ¼ T e xs : ð5Þ
To linearize (5) as in [36], one can get P e ¼ Iopt xs þ W opt iqr ,where Iopt and W opt are the appropriate constant values for cal-
culated wind power access the real values.

2.2. DFIG-based interconnected wind power systems

The LFC model of DFIG-based interconnected wind power systems can be described as
8     
>
> Df_ i ðtÞ ¼ T1pi kpi DPmi ðtÞ  DP di ðtÞ  DP ij ðtÞ  X 3i W opti Diqri ðtÞ þ Iopti Dxi ðtÞ  Df i ðtÞ
>
>
>
>
> _
> DP mi ðtÞ ¼ T chi ðDPv i ðtÞ  DPmi ðtÞÞ
> 1
>
>
> _
>
>
> DE ðtÞ ¼ ki ðDP12 ðtÞ þ Bi Df i ðtÞÞ
< i  
DP_ v i ðtÞ ¼ T1gi DPv i ðtÞ  DEi ðt  si Þ  DfRi ðtÞ þ D P ci ðtÞ ð6Þ
>
> i
>
> _
>
> DP 12 ðtÞ ¼ 2pT 1 ðDf 1 ðtÞ  Df 2 ðtÞÞ
>
>
> _
>  
>
> Di ðtÞ ¼ T11i X 2i DV qri ðtÞ  Diqri ðtÞ
>
> qri
>
: _  
Dxi ðtÞ ¼ 2M1 DT mi ðtÞ  X 3i Diqri ðtÞ
ti

with DP12 ðtÞ ¼ DP 21 ðtÞ; T pi ¼ M


Di
i
; kpi ¼ D1i ; Bi ¼ R2ii þ Di ; i; j ¼ 1; 2 and i – j. The parameters used in DFIG-based interconnected
wind power systems are given in Table 1. Fig. 1 shows the schematic diagram for the DFIG-based interconnected wind power
systems. Now, defining a state vector
h iT
xðtÞ ¼ Df 1 ðtÞ DPTm1 ðtÞ DPTv 1 ðtÞ DET1 ðtÞ Diqr1 ðtÞ DxT1 ðtÞ DPT12 ðtÞDf 2 ðtÞ DPTm2 ðtÞ DPTv 2 ðtÞ DET2 ðtÞ Diqr2 ðtÞ DxT2 ðtÞ ;
T T T T

T
the disturbance vector wðtÞ ¼ ½DP Td1 ðtÞ DT Tm1 ðtÞ DP Td2 ðtÞ DT Tm2 ðtÞ ; the control input
T
uðtÞ ¼ ½DPTc1 ðtÞ DV Tqr1 ðtÞ DPTc2 ðtÞ DV Tqr2 ðtÞ and, the output vector
T
Dx x
T T
yðtÞ ¼ ½Df 1 ðtÞ 1 ðtÞ
T
Df 2 ðtÞ D T2 ðtÞ , the state-space form of (6) is given as follows:

75
S. Kuppusamy, Young Hoon Joo and Han Sol Kim Information Sciences 582 (2022) 73–88

Table 1
Parameters of DFIG-based interconnected wind power systems [24].

DP v i Governor valve position

Df i Deviations of frequency
DP 12 Tie-line power flow from Area 1 to Area 2
DP mi Mechanical output of generator
DEi Area control error signals
DP di Load disturbances
Mi Moments of inertia of the generator
DT mi Mechanical torque variation
Mti Inertia constant of wind turbine
DP ci Load reference deviation
DP ei Output wind power
Ri Speed droops
kpi Proportional gains of local PI controllers
T gi Time constant of the governor
ki Integral gains of local PI controllers
si Constant time delay
Di Damping coefficient of generator
T chi Time constant of turbine
Bi Frequency bias parameter

Fig. 1. The fault-tolerant LFC for interconnected wind power systems.

_
xðtÞ ¼ AxðtÞ þ Bs1 xðt  s1 Þ þ Bs2 xðt  s2 Þ þ BuðtÞ þ DwðtÞ ð7Þ
yðtÞ ¼ CxðtÞ;
where
2 3 2 3
~1
A 1
A 066 1
B 072  
6  7 6 033 s
B 039
A ¼ 4 A3 0 
A3 5; B ¼ 4 072 2 7
B 5; Bs1 ¼ 1
;
0103 0101 0109
066 2
A ~2
A 012 012
2 3
   0710 071 072   
D1 072 6 s 7 C1 072
D¼ ; Bs2 ¼ 4 0310 B 032 5; C ¼ ;
072 2
D 2
072 2
C
0310 031 032
with for all i ¼ 1; 2

76
S. Kuppusamy, Young Hoon Joo and Han Sol Kim Information Sciences 582 (2022) 73–88

2 kpi kpi X 3i W opti kpi X 3i Iopti


3
 T1 T pi
0 0  T pi
 T pi
6 pi
7
6 0  T1 1
0 0 0 7
6 7
6 chi T chi
7
6 1 0  T1 0 0 0 7 h i
~i ¼ 6
A 6 Ri T gi gi
7 T
7; Bsi ¼ 0 0  T gi ;
1
6 kB 0 0 0 0 0 7
6 i i 7
6 7
6 0 0 0 0  T1 0 7
4 1i 5
X 3i
0 0 0 0  2M ti
0
2 3 2 3
1 k
0 0 T gi
0 0  Tpipi 0 0 0 0 0
T ¼ 4
B T ¼ 4
5; D 5;
i X 2i i 1
0 0 0 0 T 1i 0 0 0 0 0 2Mti

  h i
1 0 0 0 0 0
T ¼
C  3 ¼ ½ 2pT 1
;A  T ¼ ð1Þi kpi
015 ; A 0 0 ki 0 0 :
i i T pi
0 0 0 0 0 1

Remark 1. When analyzing the stability of interconnected wind power systems, the traditional LFC for each control area
with a specific integral gain does not offer stable performance or better damping responses as the delay size increases. It can
be noted from [37] that the optimal tuning of traditional PI control gain in the existence of communication delay and load
changes in respective control areas is challenging work. Hence, inspired by the works [24,25,35], the main aim of the present
study is to introduce an additional desirable control input uðtÞ instead of tuning the PI control gain that can stabilize the
proposed model (10) and to assure the H1 performance.

In the sequel, we design a fault-tolerant LFC to stabilize the DFIG-based interconnected wind power systems (7). Accord-
ing to this, let uF ðtÞ be the signal from the actuator that has failed and it will be modeled as follows:

uF ðtÞ ¼ NuðtÞ; ð8Þ


 6d 6d
where N ¼ diagfd1 ; d2 ; d3 ; d4 g be the actuator fault matrix with 0 6 d ^ ; ðl ¼ 1; 2; 3; 4Þ. Suppose, N is an unknown
l l l
 ^
actuator fault matrix, dl is an unknown constant; dl and dl are the known lower and upper bounds of dl . Now, setting
^  ^ 
N0 ¼ diagfd10 ; d20 ; d30 ; d40 g and N1 ¼ diagfd11 ; d21 ; d31 ; d41 g, where di0 ¼ di þ2 di and di1 ¼ di 2 di . Then, the actuator fault matrix
can be expressed as
N ¼ N0 þ N1 DN
where DN ¼ diagfs1 ; s2 ; s3 ; s4 g with 1 6 sl 6 1 for all l ¼ 1; 2; 3; 4. Hence, the control input can be designed in the following
form:
uðtÞ ¼ KxðtÞ þ Lxðt  qÞ ð9Þ

where q is the constant delay intensionally introduced in the controller; K and L 2 R represent the control gain matrices 413

respectively, corresponding to the present and delayed states. Hence, the dynamics of DFIG-based interconnected wind
power systems (7) with fault-tolerant LFC (8) and (9) can be formulated as

_
xðtÞ ¼ ðA þ BNKÞxðtÞ þ Bs1 xðt  s1 Þ þ Bs2 xðt  s2 Þ þ BNLxðt  qÞ þ DwðtÞ
ð10Þ
yðtÞ ¼ CxðtÞ:

Lemma 1 [38]. For all continuously differentiable function v : ½a; b ! Rn , the inequality

Z b Z b Z b
ðb  aÞ vT ðsÞU vðsÞds P vT ðsÞdsU vðsÞds þ 3/T U/
a a a
Rb Rb Rh
where / ¼ a
vðsÞds  ba
2
a b
vðsÞdsdh; holds for any matrix U > 0.

3. Design of fault-tolerant LFC scheme for DFIG-based interconnected wind power systems

The main objective of this section is to derive the stability and stabilization conditions for DFIG-based interconnected
wind power systems (10) with known and unknown actuator faults and it can be stated in the following problems:
77
S. Kuppusamy, Young Hoon Joo and Han Sol Kim Information Sciences 582 (2022) 73–88

Problem 1. Given system matrices of (10) with well known fault-tolerant LFC gains K and L, the system (10) is
asymptotically stable under the load disturbances wðtÞ which satisfies the following objectives.

 The asymptotic stability condition is obtained for proposed system without load disturbances
 For any nonzero wðtÞ 2 L2 ½0; þ1Þ with a given scalar c, the inequality
kyðtÞk2 6 ckwðtÞk2
holds under the zero initial condition.

3.1. Asymptotic stability condition under known actuator fault

In this subsection, the following theorem is derived for an asymptotic stability condition of (10), which is a solution to the
Problem 1:

Theorem 1. For given scalars sj ; q, known actuator fault matrix N and gain matrices K; L, if there are P > 0; Q j > 0; U j > 0; Sj > 0
ðj ¼ 1; 2Þ any matrices X l ðl ¼ 1; 2; 3Þ such that
X < 0; ð11Þ
where
!
n o X
2
X ¼ Sym e1 Pe5 þ e1
T
ðQ j þ ð1Þ ðjþ1Þ
Sj þ sj U j Þ þ C C eT1  e2 Q 1 eT2  e3 ðS2 þ Q 2 ÞeT3 þ s22 e5 S2 eT5  e4 S1 eT4
T

j¼1

X    T

2
1 3 2 2
þ Sym e1 S2 eT3 þ  e2jþ4 U j eT2jþ4  e2jþ4  e2jþ5 U j e2jþ4  e2jþ5
j¼1
sj sj s1 s1
nh ih io
þ Sym e5 X 1 þ e4 X 2 þ e3 X 3 eT5 þ ð A þ BNK ÞeT1 þ Bs1 eT2 þ Bs2 eT3 þ BNLeT4 þ DeT10  e10 c2 IeT10 ;

then the system (10) is asymptotically stable with H1 norm bound c.

Proof. Consider the following LKF:


X
3
VðtÞ ¼ V i ðtÞ; ð12Þ
i¼1

where
2 h
X i
Rt
V 1 ðtÞ ¼ xT ðtÞPxðtÞ þ tsj
xT ðsÞQ j xðsÞds ;
j¼1
P 2 hR 0 R t i
V 2 ðtÞ ¼ j¼1 xT ðsÞU j xðsÞdsdh ;
sj tþh
Rt R0 Rt
V 3 ðtÞ ¼ tq xT ðsÞS1 xðsÞds þ s2 s tþh x_ T ðsÞS2 xðsÞdsdh:
_
2

Calculating the time derivatives of (12) along the trajectories of (10), we get

V_ 1 ðtÞ ¼ 2xT ðtÞPxðtÞ


_ þ xT ðtÞ½Q 1 þ Q 2 xðtÞ  xT ðt  s1 ÞQ 1 xðt  s1 Þ  xT ðt  s2 ÞQ 2 xðt  s2 Þ ð13Þ
Z t Z t
V_ 2 ðtÞ ¼ xT ðtÞ½s1 U 1 þ s2 U 2 xðtÞ  xT ðsÞU 1 xðsÞds  xT ðsÞU 2 xðsÞds ð14Þ
ts1 ts2
Z t
V_ 3 ðtÞ ¼ xT ðtÞS1 xðtÞ  xT ðt  qÞS2 xðt  qÞ þ s22 x_ T ðtÞS2 xðtÞ
_  s2 x_ T ðsÞS2 xðsÞds:
_ ð15Þ
ts2

Based on Lemma 1, the integral inequalities in (14) can be written as


Z t T
1 3 2 2
 xT ðsÞU 1 xðsÞds 6  !T1 ðtÞU 1 !1 ðtÞ  ½!1 ðtÞ  !2 ðtÞ U 1 ½!1 ðtÞ  !2 ðtÞ; ð16Þ
ts1 s1 s1 s1 s1
Z t T
1 3 2 2
 xT ðsÞU 2 xðsÞds 6  !T3 ðtÞU 2 !3 ðtÞ  ½!3 ðtÞ  !4 ðtÞ U 2 ½!3 ðtÞ  !4 ðtÞ; ð17Þ
ts2 s2 s2 s2 s1
where

78
S. Kuppusamy, Young Hoon Joo and Han Sol Kim Information Sciences 582 (2022) 73–88

Z t Z t Z h Z t Z t Z h
!1 ðtÞ ¼ xðsÞds; !2 ðtÞ ¼ xðsÞdsdh; !3 ðtÞ ¼ xðsÞds; !4 ðtÞ ¼ xðsÞdsdh:
ts1 ts1 t ts2 ts2 t

Now, by applying Jensen’s inequality to the integral term in (15), we can obtain
Z t
 s2 x_ T ðsÞS2 xðsÞds
_ 6 xT ðtÞS2 xðtÞ þ xT ðtÞS2 xðt  s2 Þ þ xT ðt  s2 ÞS2 xðtÞ  xT ðt  s2 ÞS2 xðt  s2 Þ: ð18Þ
ts2

Additionally, based on system (10), the following equation holds for any appropriately dimensioned matrices X 1 ; X 2 , and
X3:

2½x_ T ðtÞX 1 þ xT ðt  qÞX 2 þ xT ðt  s2 ÞX 3 ½xðtÞ


_ þ ðA þ BNKÞxðtÞ þ Bs1 xðt  s1 Þ þ Bs2 xðt  s2 Þ þ BNLxðt  qÞ þ DwðtÞ ¼ 0:
ð19Þ
Combining (13)–(19), we obtain
_
VðtÞ 6 fT ðtÞXfðtÞ  yT ðtÞyðtÞ þ c2 wT ðtÞwðtÞ; ð20Þ
where
h i
fT ðtÞ ¼ xT ðtÞ xT ðt  s1 Þ xT ðt  s2 Þ xT ðt  qÞ x_ T ðtÞ !T1 ðtÞ !T2 ðtÞ !T3 ðtÞ !T4 ðtÞ wT ðtÞ :

From (20), one can see that if matrix inequality holds (11), then
_
VðtÞ 6 yT ðtÞyðtÞ þ c2 wT ðtÞwðtÞ: ð21Þ
_
Here, VðtÞ is continuous in time. Hence, integrating (21) on both sides from 0 to þ1, which yields
Z þ1
Vðþ1Þ  Vð0Þ 6 ½yT ðtÞyðtÞ þ c2 wT ðtÞwðtÞdt: ð22Þ
0

Under the zero initial condition Vð0Þ ¼ 0, we get


Z þ1 Z þ1
yT ðtÞyðtÞdt 6 c2 wT ðtÞwðtÞdt:
0 0

Thus, kyðtÞk2 6 ckwðtÞk2 for any nonzero wðtÞ 2 L2 ½0; 1Þ. For the condition that wðtÞ ¼ 0, there exists a scalar k > 0 such
_
that VðtÞ 6 kkxðtÞk2 for xðtÞ – 0.Hence, we can confirm that the system (10) is asymptotically stable with H1 performance.
This completes the proof of Theorem 1.

Problem 2. The fault-tolerant LFC gain matrices K and L are calculated from the solvable LMIs to ensure the stabilization
criterion for system (10) which satisfies the objectives of Problem 1.

3.2. Fault-tolerant LFC Design with Known Actuator Fault

In this subsection, we will design the control gain matrices of the fault-tolerant LFC scheme (8) for the proposed system
(10). Define X 1 ¼ Y 1 ; X 2 ¼ aY 1 ; X 3 ¼ bY 1 , and setting M ¼ KY T ; N ¼ LY T ; P ^ j ¼ YQ Y T ; ^
^ j ¼ YP j Y T ; Q ^ j ¼ YU j Y T , pre
Sj ¼ YSj Y T ; U
j
and post multiplying (11) by diagfY; Y; Y; Y; Y; Y; Y; Y; Y; Ig and its transpose, respectively. Then, based on the Schur comple-
ment, one can get the following LMI
H < 0; ð23Þ
where
" #
^
X e1 YC T
H¼ ;
 I
!
n o 2 
X    n o
^ ¼ Sym e1 Pe
X ^ T þ e1 ^ j þ ð1Þðjþ1Þ ^Sj þ sj U
Q ^j eT1  e2 Q ^ 2 eT þ s2 e5 ^S2 eT  e4 ^S1 eT þ Sym e1 ^S2 eT
^ 1 eT  e3 ^S2 þ Q
5 2 3 2 5 4 3
j¼1

X
2    T !
1 ^ 3 2 ^ 2
þ  e2jþ4 U j e2jþ4 
T
e2jþ4  e2jþ5 U j e2jþ4  e2jþ5
j¼1
sj sj s1 s1
nh ih   io
þ Sym e5 þ ae4 þ be3 Y T eT5 þ AY T þ BNM eT1 þ Bs1 Y T eT2 þ Bs2 Y T eT3 þ BNNeT4 þ DeT10  e10 c2 IeT10 :

79
S. Kuppusamy, Young Hoon Joo and Han Sol Kim Information Sciences 582 (2022) 73–88

From the above analysis, we can conclude the following theorem to design the control gains K and L as a solution to Prob-
lem 2.

^ j > 0; U
^ > 0; Q
Theorem 2. For given scalars sj ; q, known actuator fault matrix N, if there are P ^ j > 0; ^
Sj > 0 ðj ¼ 1; 2Þ any matrices
M and N such that the LMI (23) holds, then the system (10) is asymptotically stable under the control input (9) with H1 norm
bound c. Moreover, the control gain matrices are obtained as K ¼ MY T and L ¼ NY T .

3.3. Fault-tolerant LFC design with unknown actuator fault

In this subsection, we will discuss the asymptotic stability of interconnected wind power systems under an unknown or
unfixed actuator fault matrix.

^ > 0; U
^ > 0; Q
Theorem 3. For given scalars sj ; q; a, and b, unknown actuator fault matrix N, if there are P ^ j > 0; ^
Sj > 0
j
ðj ¼ 1; 2Þ any matrices Y; M; N and scalar  > 0 such that the following LMI holds
2 3
H1 X2 XT3
6 7
4  I 0 5 < 0; ð24Þ
  I

where X2 ¼ ½M 0 0 N 0 0 0 0 0 0 0T ; X3 ¼ ½0 0 bNT1 BT aNT1 BT NT1 BT 0 0 0 0 0 0 and H1 can be obtained from H in (23) by replac-
ing N with N0 , then the system (10) is asymptotically stable under the control input (9) with H1 norm bound c. Furthermore,
the control gain matrices are obtained as K ¼ MY T and L ¼ NY T .

Proof. The proof can be derived from Theorem 2 by replacing N with N0 þ N1 DN in (23), then we get

H1 þ X2 DNX3 þ XT3 DNT XT2 < 0:


By applying the lemma as in [39] (Lemma 2.6), we have

H1 þ 1 X2 XT2 þ XT3 X3 < 0:


Further, by applying Schur complement, one can obtain (24). This proof is completed.

4. Design of H ‘ LFC scheme for interconnected power systems

In this section, we consider the interconnected delayed power systems without DFIG-based wind turbines; that is, we
focus only on the power systems model, which is a special case of previous section. The dynamics of the interconnected
power systems with time delays can be described for i; j ¼ 1; 2 ði – jÞ as follows:
8    
>
>
> Df_ i ðtÞ ¼ T1pi kpi DPmi ðtÞ  DP di ðtÞ  DP ij ðtÞ  Df i ðtÞ
>
>
>
> _
> DP mi ðtÞ ¼ T chi ðDPv i ðtÞ  DPmi ðtÞÞ
> 1
<
> DE_ i ðtÞ ¼ ki ðDP12 ðtÞ þ Bi Df i ðtÞÞ ð25Þ
>
>  
>
> DP_ v i ðtÞ ¼ T1gi DPv i ðtÞ  DEi ðt  si Þ  DfRi ðtÞ þ DPci ðtÞ
>
>
>
>
i
: _
DP 12 ðtÞ ¼ 2pT 1 ðDf 1 ðtÞ  Df 2 ðtÞÞ
T
with DP12 ðtÞ ¼ DP21 ðtÞ. Now, taking a state vector xðtÞ ¼ ½Df 1 ðtÞ DP Tm1 ðtÞ DP Tv 1 ðtÞDET1 ðtÞ DPT12 ðtÞ
T
Df 2 ðtÞ DP Tm2 ðtÞ T
DPv 2 ðtÞ DET2 ðtÞT , the load disturbance vector
T T
wðtÞ ¼ ½DP Td1 ðtÞ DP Td2 ðtÞ , control input uðtÞ ¼ ½DP Tc1 ðtÞ DPTc2 ðtÞ , then the state space model of interconnected power sys-
tems can be formulated as follows:
_
xðtÞ ¼ AxðtÞ þ Ad1 xðt  s1 Þ þ Ad2 xðt  s2 Þ þ BuðtÞ þ DwðtÞ ð26Þ
yðtÞ ¼ CxðtÞ;
and the constant matrices A; B; C; D; Ad1 ; Ad2 are same as in [24] and it is omitted there. To obtain the asymptotic stability and
stabilization conditions for system (26), the desired control law uðtÞ is designed as
uðtÞ ¼ KxðtÞ þ Lxðt  qÞ ð27Þ

80
S. Kuppusamy, Young Hoon Joo and Han Sol Kim Information Sciences 582 (2022) 73–88

where K 2 R29 and L 2 R29 are control gain matrices. Now, the sufficient conditions are derived for asymptotic stability of
(26) under the designed controller (27), and minimize the H1 performance index c, with the objectives satisfied as in Prob-
lem 1.

4.1. Delay-Dependent H1 LFC Design

In this subsection, the asymptotic stability condition for interconnected power systems is derived under the present and
delayed state feedback control.

^ j > 0; U
^ > 0; Q
Theorem 4. For given scalars sj ; q; a, and b, if there are P ^ j > 0; ^
Sj > 0 ðj ¼ 1; 2Þ any matrices Y; M and N such that
the following LMI holds
" #
~ e1 YC T
X
; ð28Þ
 I

where
!
 n o 2 
X    n o
^
X ¼ Sym e1 Pe4 þ e1
T ^
Q j þ ð1Þ ðjþ1Þ ^ ^
Sj þ sj U j eT1  e2 Q ^ 2 eT þ s2 e5 ^S2 eT  e4 ^S1 eT þ Sym e1 ^S2 eT
^ 1 eT  e3 ^S2 þ Q
2 3 2 5 4 3
j¼1

X
2    T !
1 ^ j eT  3 2 ^ j e2jþ4  2 e2jþ5
þ  e2jþ4 U e2jþ4  e2jþ5 U
j¼1
sj 2jþ4
sj s1 s1
nh ih   io
þ Sym e5 þ ae4 þ be3 Y T eT5 þ AY T þ BM eT1 þ Ad1 Y T eT2 þ Ad2 Y T eT3 þ BNeT4 þ DeT10  e10 c2 IeT10 :

Then, the system (26) is asymptotically stable under the control input (27) with H1 norm bound c. Moreover, the control gain
matrices are obtained as K ¼ MY T and L ¼ NY T .

Proof. The proof of this theorem is immediately obtained from Theorem 2.

4.2. State Feedback H1 LFC Design

In this subsection, the delay-dependent stability condition for interconnected power systems (26) under the state feed-
back control law without the delayed state, that is uðtÞ ¼ KxðtÞ.

^ j > 0; U
^ > 0; Q
Theorem 5. For given scalars sj ; a, and b, if there are P ^ j > 0; ^
S2 > 0 ðj ¼ 1; 2Þ any matrices Y; M and N such that the
following LMI holds
" #
~ e1 YC T
X
; ð29Þ
 I
 n o P      n o P
where ^ T þ e1
X ¼ Sym e1 Pe 2
Q ^j  ^
^ j þ sj U ^ 1 eT  e3 ^
S2 eT1  e2 Q ^ 2 eT þ s2 e4 ^
S2 þ Q S2 eT4 þ Sym e1 ^
S2 eT3 þ 2j¼1
5 j¼1 2 3 2
h i h iT nh ih   io
^ j eT  3 e2jþ3  2 e2jþ4 U
 s1 e2jþ3 U ^ j e2jþ3  2 e2jþ4 T
þ Sym e4 þ be3 Y T eT5 þ AY þ BM eT1 þ Ad1 Y T eT2 þ Ad2 Y T eT3 þ DeT9 
j 2jþ3 sj s1 s1

e9 c2 IeT9 . Then, the system (26) is asymptotically stable under the state feedback control law uðtÞ ¼ KxðtÞ with H1 norm bound c. Moreover,
the control gain matrices are obtained as K ¼ MY T .

Proof. Applying the same procedure as in Theorem 2 by replacing with S1 ¼ L ¼ 0; q ¼ 0; N ¼ I, and


h i
fT ðtÞ ¼ xT ðtÞ xT ðt  s1 Þ xT ðt  s2 Þ x_ T ðtÞ !T1 ðtÞ !T2 ðtÞ !T3 ðtÞ !T4 ðtÞ wT ðtÞ , we obtain (29) which completes the proof
of this theorem.

Remark 2. The main contributions of this work are listed as follows:

1. Unlike the previous studies, the delayed interconnected power systems are adapted together with DFIG-based wind
farms.
2. To attenuate the load disturbances and wind speed fluctuations for DFIG-based interconnected wind power systems, the
fault-tolerant H1 LFC is designed for both cases of known and unknown actuator faults.

81
S. Kuppusamy, Young Hoon Joo and Han Sol Kim Information Sciences 582 (2022) 73–88

3. To strengthen the main result of this work, two types of H1 LFC, with and without constant time-delay, are designed for
interconnected power systems with communication delays.
4. When compared to [24], the superiority of proposed method will be given in the following numerical examples via
obtaining the minimum H1 norm bound.

5. Numerical examples

In this section, simulation results are provided to illustrate the validity and superiority of the proposed control design
method.

Example 1. In this example, the corresponding nominal parameters of the DFIG-based interconnected wind power systems
are given in p.u., and listed in Table 2. With the above system parameters, we chosen the constant-time delays
s1 ¼ 0:1; s2 ¼ 0:6 and q ¼ 0:5; and scalars are a ¼ 0:0015; b ¼ 0:7.

5.1. Actuator fault is known

In this subsection, the known actuator fault matrix is selected as N ¼ diagf0:5; 0:7; 0:3; 0:1g. Furthermore, the disturbance
input including the load disturbances and wind speed fluctuations, is taken as

½0:01; 0:01; 0:01; 0:01p:u:; ift ¼ 10s; 20s; 30s; 40s
wðtÞ ¼
0 otherwise:
Based on Theorem 2 with the above discussion, the fault-tolerant control gains (30) and (31) are determined by solving
LMI (23).
2 3
0:0165 0:0007 0:0002 0:0004 0:0000 0:0420 0:0008 0:0024 0:0000 0:0000 0:0000 0:0000 0:0002
6 0:4406 0:0098 0:0025 0:0123 0:0178 0:1841 0:0028 0:0424 0:0006 0:0001 0:0006 0:0008 0:0201 7
46 7
K ¼ 10  6 7
4 0:0007 0:0001 0:0000 0:0011 0:0000 0:0085 0:0031 0:0746 0:0015 0:0019 0:0006 0:0003 0:2888 5
0:2105 0:0029 0:0000 0:0317 0:0031 0:0085 0:0342 2:2184 0:0331 0:0220 0:0022 0:1092 0:5613
ð30Þ

2 3
0:0009 0:0000 0:0000 0:0000 0:0000 0:0005 0:0000 0:0015 0:0000 0:0000 0:0000 0:0000 0:0004
6 0:1099 0:0016 0:0001 0:0036 0:0019 0:0238 0:0004 0:1157 0:0012 0:0000 0:0012 0:0024 0:0329 7
6 7
L¼6 7 ð31Þ
4 0:0142 0:0002 0:0000 0:0002 0:0002 0:0029 0:0008 0:0414 0:0007 0:0006 0:0000 0:0006 0:1927 5
0:5442 0:0081 0:0001 0:0165 0:0110 0:2386 0:0098 2:2062 0:0218 0:0003 0:0198 0:0451 0:7765

Also, the minimum H1 norm bound can be calculated as c ¼ 0:096. In this connection, the simulation results for the state
trajectories of system (10) are plotted in Fig. 2a and 2b. It is observed from these figures that the DFIG-based interconnected
wind power systems include the load disturbances and wind speed fluctuation, the frequency deviation can be effectively
limited within the interval ½0:12; 0:07 under the designed fault-tolerant LFC input. Hence, the designed fault-tolerant
LFC scheme guarantees the asymptotic stability of DFIG-based interconnected wind power systems under known actuator
fault. In addition to this, the trajectories of fault-tolerant LFC input are displayed in Fig. 3.

5.2. Actuator fault is unknown

In this case, the actuator fault happens in a range of interval; that is 0:1 6 d1 6 0:7; 0:4 6 d2 6 0:6; 0:7 6 d3 6 0:9, and
0:2 6 d4 6 0:8. With the same parameters considered in Subsection 5:1, by solving the LMI (24) in Theorem 3, we can obtain
the feasible solution and minimum H1 norm bound as c ¼ 0:0501. By utilizing the above, in order to confirm the utility of
the fault-tolerant LFC design, the simulation result of the DFIG-based interconnected wind power systems (10) in the pres-
ence of an unknown actuator fault are displayed in Figs. 4a and 4b. Figs. 4a and 4b reveal that the designed fault-tolerant LFC
stabilized the system (10) even if the actuator fault happens in a range of intervals. In addition, the designed controller effec-
tively restrain the frequency deviation within the interval ½0:015; 0:015 for system (10) which ensures the desired fre-
quency. Moreover, the evolution of fault-tolerant controller (9) with unknown actuator fault is presented in Fig. 5.

Example 2. In this example, the two-area power system is considered, where Area-1 and Area-2 are modeled with two and
four generators, respectively; and the parameter values are given as in [24]:

Table 2
Parameters of TAPSWF.

Parameters T chi ki Ri Di T gi Mi W Oi T 1i IOi Mti X 2i X 3i T1

Area-1 0.3 0.5 0.05 1.0 0.1 10 1.17 1382.18 0.8 9 181.16 0.98 0.2
Area-2 0.17 0.5 0.05 1.5 0.4 12 1.20 1482.18 0.9 10 200.00 1.00

82
S. Kuppusamy, Young Hoon Joo and Han Sol Kim Information Sciences 582 (2022) 73–88

Fig. 2. State trajectories for system (10) with known actuator faults (a) Area-1 (b) Area-2.

Fig. 3. Time responses of fault-tolerant control input for known actuator fault case.

Fig. 4. State trajectories for system (10) via fault-tolerant control input with unknown actuator faults (a) Area-1 (b) Area-2.

Fig. 5. Time responses of fault-tolerant control input with unknown actuator fault.

83
S. Kuppusamy, Young Hoon Joo and Han Sol Kim Information Sciences 582 (2022) 73–88

Fig. 6. Evolution of load disturbances for delayed feedback control input case (a) Area-1 (b) Area-2.

Fig. 7. Dynamic response of frequency deviations for system (26)with delayed feedback control.

Fig. 8. Dynamic response of mechanical power output of generators for system (26) with delayed feedback control.

For Area-1: T ch1 ¼ 0:3; R1 ¼ 0:05; D1 ¼ 1; T g1 ¼ 0:1; M 1 ¼ 10; T 1 ¼ 0:2, and k1 ¼ 0:5.
For Area-2: T ch2 ¼ 0:17; R2 ¼ 0:05; D2 ¼ 1:5; T g2 ¼ 0:4; M 2 ¼ 12, and k2 ¼ 0:5. The constants are taken as
b ¼ 0:07; a ¼ 0:01. The communication time-delays in Area-1 and Area-2 are selected as s1 ¼ 0:1 and s2 ¼ 0:6, respectively.

5.2.1. Under the delayed feedback control strategy


The signal transmission delay of the controller is chosen as q ¼ 0:70, then solving the LMI (28), we obtain the control
gains as follows:
 
1:2784 0:0109 0:0008 0:0039 0:0060 0:0330 0:0002 0:0000 0:0001
K ¼ 104 
0:2611 0:0020 0:0001 0:0301 0:0162 3:9763 0:0222 0:0035 0:0007
 
0:2927 0:0036 0:0004 0:0023 0:0040 0:0799 0:0003 0:0000 0:0000

0:1967 0:0004 0:0000 0:0423 0:0439 1:5575 0:0007 0:0008 0:0009

84
S. Kuppusamy, Young Hoon Joo and Han Sol Kim Information Sciences 582 (2022) 73–88

Also, the minimum H1 norm bound can be calculated as c ¼ 0:0119. With the above parameters and control gains, the
simulation results of system (26) under the control law (27) are displayed in Figs. 6–9. In Figs. 6a and 6b, the random step
changes of load demand in Area-1 and Area-2 are depicted. The deviations of frequency and the mechanical power output of
generator for both control areas under the delayed feedback controller are given in Figs. 7 and 8. In addition, Fig. 7 exhibits
the deviations of frequency could approach zero within a short period. Further, the corresponding control inputs are shown
in Figs. 9. It can be concluded from these figures, the two-area power systems with time-delays are stabilized with the min-
imum H1 norm bound c ¼ 0:0119 via the designed delayed feedback control.

5.2.2. Under the state feedback control strategy


In this case, the two-area power systems with time-delays via state feedback control law can be analyzed. Let us take the
previous case parameters value, the following control gain is obtained by solving the LMI (29):
 
1:6212 0:0087 0:0007 0:0045 0:0027 0:0522 0:0002 0:0000 0:0000
K ¼ 104  : ð32Þ
0:4750 0:0020 0:0001 0:0290 0:0268 4:9856 0:0180 0:0029 0:0012

Fig. 9. Time responses of delayed feedback control input.

Fig. 10. Time responses of load disturbances for state feedback control input case (a) Area-1 (b) Area-2.

Fig. 11. Time responses of Df i ðtÞ ði ¼ 1; 2Þ with the control gain (32).

85
S. Kuppusamy, Young Hoon Joo and Han Sol Kim Information Sciences 582 (2022) 73–88

Fig. 12. Time responses of DP mi ðtÞ ði ¼ 1; 2Þ with the state feedback control gain (32).

Fig. 13. Time responses of state feedback control input.

Table 3
Comparison results of minimum H1 norm bound c.

Control Proposed method [24]


Delayed feedback 0:0119 4:0124
State feedback 0:0047 0:4493

Also, the minimum H1 norm bound can be calculated as c ¼ 0:0047. By using the control gain (32), the simulation results
of the two-area power systems with time-delays via state feedback control law are depicted in Figs. 10–13. The random load
disturbances for each area are given in Figs. 10a and 10b. From Fig. 11, the frequency deviations of Area-1 and Area-2 are
reaches to zero within short time which shows that the designed state feedback control strategy can enhance the frequency
stability of both areas, and the designed control is well effective. Also, the mechanical power output of generator for both
control areas under the state feedback control are depicted in Fig. 12. Further, the corresponding control inputs are displayed
in Fig. 13. The Figs. 10–13 clearly shows that the designed state feedback control effectively stabilizes the two-area power
systems, including the load disturbances and communication time-delays. Moreover, to show the superiority of the proposed
method over a previous study under both delayed and state feedback control law, the comparison results of H1 norm bound
are tabulated in Table 3. Compared with the previous study [24], the minimum H1 norm bound can be obtained for two-area
power systems (26) by the proposed method.

6. Conclusion

This paper has been analyzed the fault-tolerant LFC design for DFIG-based interconnected wind power systems. The main
advantage of this paper is to design the LFC with known and unknown actuator faults to improve the frequency stability of
the DFIG-based interconnected wind power systems, including the load disturbances and wind speed fluctuations. Then, suf-
ficient conditions have been derived in the form of LMIs based on the LKF method and Wirtinger-based inequality to guar-
antee the asymptotic stability of proposed systems under the designed fault-tolerant LFC. In addition, H1 -based delayed
feedback and state feedback LFC have been designed to ensure the asymptotic stability of two-area power systems with
86
S. Kuppusamy, Young Hoon Joo and Han Sol Kim Information Sciences 582 (2022) 73–88

communication time-delays. Finally, the derived theoretical results have been validated through numerical examples. Future
research directions will include the investigation of fault-tolerant control design for DFIG-based interconnected wind power
systems with parametric uncertainties and time-varying delays.

CRediT authorship contribution statement

Subramanian Kuppusamy: Conceptualization, Methodology, Writing - original draft, Validation. Young Hoon Joo:
Supervision, Validation, Writing - review & editing. Han Sol Kim: Writing - review & editing.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have
appeared to influence the work reported in this paper.

Acknowledgements

This work was partially supported by the Basic Science Research Program through the National Research Foundation of
Korea (NRF) funded by the Ministry of Education (NRF-2016R1A6A1A03013567, NRF-2021R1A2B5B01001484) and by the
Korea Institute of Energy Technology Evaluation and Planning (KETEP) and the Ministry of Trade, Industry & Energy(MOTIE)
of the Republic of Korea (No. 20194030202300).

References

[1] M.A. Mahmud, M.J. Hossain, H.R. Pota, M.S. Ali, Generalized Lyapunov function for stability analysis of interconnected power systems, in: AUPEC, IEEE,
2011, pp. 1–6.
[2] Z. Hu, S. Liu, W. Luo, L. Wu, Resilient distributed fuzzy load frequency regulation for power systems under cross-layer random denial-of-service attacks,
IEEE Trans. Cybern. (2020), https://doi.org/10.1109/TCYB.2020.3005283.
[3] H.J. Lee, J.B. Park, Y.H. Joo, Robust load-frequency control for uncertain nonlinear power systems: A fuzzy logic approach, Inf. Sci. 176 (23) (2006) 3520–
3537.
[4] G.B. Koo, J.B. Park, Y.H. Joo, Decentralized sampled-data fuzzy observer design for nonlinear interconnected systems, IEEE Trans. Fuzzy Syst. 24 (3)
(2016) 661–674.
[5] P.M. Pardalos, S. Rebennack, M.V. Pereira, N.A. Iliadis, V. Pappu (Eds.), Handbook of wind power systems, Springer, Berlin Heidelberg, 2013.
[6] M.S. Lu, C.L. Chang, W.J. Lee, L. Wang, Combining the wind power generation system with energy storage equipment, IEEE Trans. Ind. Appl. 45 (6)
(2009) 2109–2115.
[7] L. Shanmugam, Y.H. Joo, Stabilization of permanent magnet synchronous generator-based wind turbine system via fuzzy-based sampled-data control
approach, Inf. Sci. 559 (2021) 270–285.
[8] S.M. Muyeen, H.M. Hasanien, J. Tamura, Reduction of frequency fluctuation for wind farm connected power systems by an adaptive artificial neural
network controlled energy capacitor system, IET Renew. Power Gener. 6 (4) (2012) 226–235.
[9] M. Ma, X. Liu, C. Zhang, LFC for multi-area interconnected power system concerning wind turbines based on DMPC, IET Generation, Transmission
Distribution 11 (10) (2017) 2689–2696.
[10] P. Mani, Y.H. Joo, Fuzzy logic-based integral sliding mode control of multi-area power systems integrated with wind farms, Inf. Sci. 545 (2021) 153–169.
[11] E. Tómasson, L. Söder, Generation adequacy analysis of multi-area power systems with a high share of wind power, IEEE Trans. Power Syst. 33 (4)
(2017) 3854–3862.
[12] V.T. Phan, H.H. Lee, Improved predictive current control for unbalanced stand-alone doubly-fed induction generator-based wind power systems, IET
Electr. Power Appl. 5 (3) (2011) 275–287.
[13] M.A. Mohamed, A.A.Z. Diab, H. Rezk, T. Jin, A novel adaptive model predictive controller for load frequency control of power systems integrated with
DFIG wind turbines, Neural Comput. Appl. (2019), https://doi.org/10.1007/s00521-019-04205-w.
[14] Z.S. Zhang, Y.Z. Sun, J. Lin, G.J. Li, Coordinated frequency regulation by doubly fed induction generator-based wind power plants, IET Renew. Power
Gener. 6 (1) (2012) 38–47.
[15] Y. Liu, Q.H. Wu, X.X. Zhou, Co-ordinated multiloop switching control of DFIG for resilience enhancement of wind power penetrated power systems,
IEEE Trans. Sustainable Energy 7 (3) (2016) 1089–1099.
[16] M. Toulabi, S. Bahrami, A.M. Ranjbar, An input-to-state stability approach to inertial frequency response analysis of doubly-fed induction generator-
based wind turbines, IEEE Trans. Energy Convers. 32 (4) (2017) 1418–1431.
[17] L. Jiang, W. Yao, Q.H. Wu, J.Y. Wen, S.J. Cheng, Delay-dependent stability for load frequency control with constant and time-varying delays, IEEE Trans.
Power Syst. 27 (2) (2011) 932–941.
[18] Y. Mi, Y. Fu, D. Li, C. Wang, P.C. Loh, P. Wang, The sliding mode load frequency control for hybrid power system based on disturbance observer, Int. J.
Electrical Power Energy Systems 74 (2016) 446–452.
[19] J. Liu, Y. Gu, L. Zha, Y. Liu, J. Cao, Event-triggered H1 load frequency control for multiarea power systems under hybrid cyber attacks, IEEE Trans. Syst.,
Man, Cybern.: Syst. 49 (8) (2019) 1665–1678.
[20] C. Peng, J. Zhang, H. Yan, Adaptive event-triggering H1 load frequency control for network-based power systems, IEEE Trans. Industr. Electron. 65 (2)
(2017) 1685–1694.
[21] E. Tian, C. Peng, Memory-based event-triggering H_1) load frequency control for power systems under deception attacks, IEEE Trans. Cybern. (2020),
https://doi.org/10.1109/TCYB.2020.2972384.
[22] Y. Mi, X. Hao, Y. Liu, Y. Fu, C. Wang, P. Wang, P.C. Loh, Sliding mode load frequency control for multi-area time-delay power system with wind power
integration, IET Generation, Transmission & Distribution 11 (18) (2017) 4644–4653.
[23] D. Qian, S. Tong, H. Liu, X. Liu, Load frequency control by neural-network-based integral sliding mode for nonlinear power systems with wind turbines,
Neurocomputing 173 (2016) 875–885.
[24] R. Dey, S. Ghosh, G. Ray, A. Rakshit, H_1) load frequency control of interconnected power systems with communication delays, Int. J. Electrical Power
Energy Systems 42 (1) (2012) 672–684.
[25] Y. Sun, N. Li, X. Zhao, Z. Wei, G. Sun, C. Huang, Robust H1 load frequency control of delayed multi-area power system with stochastic disturbances,
Neurocomputing 193 (2016) 58–67.

87
S. Kuppusamy, Young Hoon Joo and Han Sol Kim Information Sciences 582 (2022) 73–88

[26] A. Ahmadi, M. Aldeen, An LMI approach to the design of robust delay-dependent overlapping load frequency control of uncertain power systems, Int. J.
Electrical Power Energy Systems 81 (2016) 48–63.
[27] S. Arunagirinathan, P. Muthukumar, Y.H. Joo, Robust reliable H1 control design for networked control systems with nonlinear actuator faults and
randomly missing measurements, Int. J. Robust Nonlinear Control 29 (12) (2019) 4168–4190.
[28] D. Rotondo, M. Buciakowski, M. Witczak, Simultaneous state and process fault estimation in linear parameter varying systems using robust quadratic
parameter varying observers, Int. J. Robust Nonlinear Control (2021), https://doi.org/10.1002/rnc.5395.
[29] S. Kuppusamy, Y.H. Joo, Stabilization of interval type-2 fuzzy-based reliable sampled-data control systems, IEEE Trans. Cybern. (2020), https://doi.org/
10.1109/TCYB.2020.3001609.
[30] J. Liu, D. Yue, Event-triggering in networked systems with probabilistic sensor and actuator faults, Inf. Sci. 240 (2013) 145–160.
[31] V. Sharmila, R. Rakkiyappan, Y.H. Joo, Fuzzy sampled-data control for DFIG-based wind turbine with stochastic actuator failures, IEEE Trans. Systems,
Man, Cybern.: Systems (2019), https://doi.org/10.1109/TSMC.2019.2946873.
[32] X. Xie, D. Yue, J.H. Park, Observer-based fault estimation for discrete-time nonlinear systems and its application: A weighted switching approach, IEEE
Trans. Circuits Syst. I Regul. Pap. 66 (11) (2019) 4377–4387.
[33] H. Yang, S. Yin, O. Kaynak, Neural network-based adaptive fault-tolerant control for Markovian jump systems with nonlinearity and actuator faults,
IEEE Trans. Systems, Man, Cybern.: Systems (2020), https://doi.org/10.1109/TSMC.2020.3004659.
[34] H. Yang, Y. Jiang, S. Yin, Adaptive fuzzy fault tolerant control for Markov jump systems with additive and multiplicative actuator faults, IEEE Trans.
Fuzzy Syst. (2020).
[35] D. Aravindh, R. Sakthivel, and S. Marshal Anthoni, Extended dissipativity-based non-fragile control for multi-area power systems with actuator fault,
International Journal of Systems Science, 50 (2)(2019) 256-272..
[36] F. Liu, J. Ma, Equivalent input disturbance-based robust LFC strategy for power system with wind farms, IET Generation, Transmission Distribution 12
(20) (2018) 4582–4588.
[37] H. Bevrani, T. Hiyama, Robust decentralized PI based LFC design for time-delay power system, Energy Convers Manage 49 (2008) 193–204.
[38] A. Seuret, F. Gouaisbaut, Wirtinger-based integral inequality: Application to time-delay systems, Automatica 49 (9) (2013) 2860–2866.
[39] R. Sakthivel, C. Wang, S. Santra, B. Kaviarasan, Non-fragile reliable sampled-data controller for nonlinear switched time-varying systems, Nonlinear
Analysis: Hybrid Systems 27 (2018) 62–76.

88

You might also like