Next Article in Journal
Optimizing the Maximum Lyapunov Exponent of Fractional Order Chaotic Spherical System by Evolutionary Algorithms
Previous Article in Journal
A Theoretical and Numerical Study on Fractional Order Biological Models with Caputo Fabrizio Derivative
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Integer and Fractional-Order Sliding Mode Control Schemes in Wind Energy Conversion Systems: Comprehensive Review, Comparison, and Technical Insight

1
School of Electrical and Electronics Engineering, Chung-Ang University, Dongjak-gu, Seoul 06974, Korea
2
Department of Electrical and Computer Engineering, COMSATS University Islamabad, Abbottabad Campus, Abbottabad 22060, Pakistan
3
Department of Electrical Engineering, University of Engineering and Technology Peshawar, Bannu Campus, Bannu 28100, Pakistan
4
Department of Electrical and Computer Engineering, COMSTAS University Islamabad, Islamabad 45550, Pakistan
5
Advanced Power and Energy Center, EECS Department, Khalifa University, Abu Dhabi P.O. Box 127788, United Arab Emirates
6
Department of Intelligent Energy and Industry, Chung-Ang University, Dongjak-gu, Seoul 06974, Korea
*
Author to whom correspondence should be addressed.
Fractal Fract. 2022, 6(8), 447; https://doi.org/10.3390/fractalfract6080447
Submission received: 13 June 2022 / Revised: 10 August 2022 / Accepted: 12 August 2022 / Published: 17 August 2022

Abstract

:
The technological development in wind energy conversion systems (WECSs) places emphasis on the injection of wind power into the grid in a smoother and robust way. Sliding mode control (SMC) has proven to be a popular solution for the grid-connected WECS due to its robust nature. This paper reviews the enhancement trends in the integer-order SMC (IOSMC) and fractional-order SMC (FOSMC) schemes reported in reputed journals over the last two decades. This work starts with a mathematical description of the wind turbine, generators, grid, and SMC and its variants available in literature. A comprehensive literature review is tabulated that includes the proposed errors, sliding surfaces, typologies, and major outcomes. Moreover, a comparative analysis of the integer-order and fractional-order SMC and its variants is also presented in this paper. This paper will provide insight for the researcher working in the WECS and will serve them in the selection and exploration of the most appropriate control schemes for quality wind power extraction. The concise mathematical proofs of the IOSMC, FOSMC and their variants will also serve the researchers in selecting the relevant sliding surfaces control laws for their research tasks. This paper also provides a comparative analysis of IOSMC, FOSMC, and fuzzy-FOSMC in terms of chattering reduction, robustness, and computational complexities using mathematical theories, simulation carried out in Matlab/Simulink, and a processor in the loop (PIL)-based experimental environment.

1. Introduction

The trend in the consumption of fossil fuels for power generation is declining with the advent of renewable-energy-generating systems. This is evident from the declining trend of fossil fuels from 86% in 1973 to 81% in 2016 [1]. Wind energy is sharing a substantial part in the overall renewable energy generation all over the world. The wind energy is utilized through a wind energy conversion system (WECS) that consists of three major components, including a wind turbine, gear, and electrical generator. The WECS converts the wind energy into electricity, and then power electronics come into play to integrate the electric power to the grid [2]. There is rapid growth in the wind turbine rotor size; thus, the WECS power-generating capacity is also progressing. The increasing trend of the rated energy output of wind turbines with the increase in rotor diameter with time is elaborated in [3]. The WECS is becoming smaller in size and less costly due to the rapid development in the power converters [4,5,6,7]. The WECS is complex and depends on external meteorological conditions, such as wind speed, and internal nonlinear dynamics, parametric uncertainties, and unknown disturbances. Thus, a control system is essential in order to transfer wind energy to the grid [8]. Various types of mechanical and electrical control techniques employed in WECS are shown in Figure 1. The WECS works in four regions of operation. In the first region ( R 1 ), the generator’s wind speed is low enough to produce power. The second region ( R 2 ) is bounded by wind speeds V i n and V r a t e d , where V r a t e d provides the rated power P r a t e d . The third region ( R 3 ) is bounded by the wind speeds V r a t e d and V o u t , and the rated power is extracted from the wind speeds in this region. The turbine is stopped for the wind speeds exceeding V o u t and is called the fourth region ( R 4 ).The four operating regions of WECS are illustrated in Figure 2 [9,10,11]. The maximum output power extraction from region R 2 is carried out through algorithms known as maximum power point tracking (MPPT) algorithms. These MPPT algorithms are only employed for variable speed wind turbines (VSWT) to extract maximum power from the wind with full or partial-order power electronics converter [8,12,13,14]. The maximization of extracted power is preserved by the VSWTs using the optimum tip speed ratio (OTSR). Extensive research has been carried out to formulate control techniques and maximum power point tracking (MPPT) in WECS.
Various typologies and configurations are developed utilizing the MPPT techniques to improve power conversion technologies and employing different generator types [15]. For instance, DC generators, synchronous generators (SGs), and induction generators (IGs) are the three major types of generators used in WECS that operate using wound rotors [16,17]. Other types of generators reported in the literature employed in WECS are wound rotor induction generators (WRIGs) [18], squirrel cage induction generators (SCIG) [19,20,21], DC generators [17], variable reluctance generators (VRGs) [22], switched reluctance generator (SRGs) [23], and permanent magnet synchronous generators (PMSGs) [24,25]. Two major categories of WRIGs are DFIGs and dynamic slip-controlled generators [26]. These generators are compared regarding their advantages and disadvantages in Figure 3. Electronic devices are the backbone for the control and flexible operation of grid-connected WECS in transient and steady-state operation [27,28]. Before introducing AC-DC-AC converters in the 2000s, soft starters were initially used in the 1980s for the grid connection of SCIGs [29]. In contrast, chopper circuits combined with diode bridges were utilized to control WRIGs using rotor resistance [30]. The back-to-back converters have the advantages of transferring a high and smooth electricity quality after extracting maximum power from wind [31,32]. In grid-connected operation, the DFIGs use partially scaled [33,34] back-to-back converters rated at 30% of the generator rating [35] and have the advantage of speed control using AC excitation [36]. The stator of DFIG is directly connected to the grid, whereas the rotor is connected to the grid using back-to-back converters. The partially scaled converters control the rotor speed and perform active and reactive power decoupled control with reactive power compensation [37]. The disadvantage of DFIG-based WECS is the protection problems under grid fault perturbations and the use of slip rings.
In contrast, the PMSG-WECS employs full-scale power converters and thus provides a complete decoupling from grid faults, which is much better [38,39]. Several typologies of grid-connected direct-driven PMSG-WECS are presented in [29,40,41,42,43], which include diode rectifiers [44], two-level back-to-back converters [45], Z-sources, multilevel converters, matrix converters, and nine switch ac-ac converter typologies. Multipole wound rotor synchronous generators are commonly used in low or variable speed areas due to their simple pitch control, low cost, high mechanical stress absorption, and torque ripple compensation capability [33]. SCIGs also provide reliable and straightforward applications in grid-connected WECS due to their ability to limit the output power; thus, they are used in high-power security and maintenance-requiring situations [46]. The SCIG generators are employed for offshore wind farms where the electricity is transferred using HVDC transmission lines. The connection of various SCIGs is established through a thyristor soft starter that reduces the mechanical stress and limits the inrush currents. The DC generators are rarely used generators in WECS due to their high maintenance and cost. The DC generators are usually employed with the small turbines together with battery banks [47,48].

2. Motivation of This Study

The MPPT techniques presented in the literature used to transfer the power of wind energy to the grid side are reported in [8]. The various MPPT control techniques prominent in the literature are OTSR control [49,50,51,52,53], WT power curves-based control [54], power signal feedback control (PSF) [8,49,55], generator signal feedback control [54], optimal torque algorithm (OTA) [50,53,56], speed sensorless control [54], and soft computing, e.g., AI and FLC-based MPPT control schemes [57,58,59,60,61,62,63]. The MPPT techniques are categorized into two groups based on power maximization, which are: (1) indirect power control (IPC) and (2) direct power control (DPC). In DPC, the electrical power ( P e ) is maximized, whereas, in IPC, the mechanical power ( P m ) is maximized. The relation between ( P m ) and ( P e ) depicts that the maximum power cannot be guaranteed even if the optimal ( P m ) is achieved. The IPC algorithm covers various types of MPPT algorithms, which are OTSR, PSF [64], and OTA [8,64,65]. The OTSR is the turbine-rotating-speed-to-wind-speed ratio, as the VSWTs are subjected to varying rotation proportional to the instantaneous change in wind speed [66]. The OTSR-based MPPT method uses an anemometer and is inefficient due to the wind speed variation along the blade length. The PSF and OTA-based methods operate without an anemometer, but these methods are still inefficient due to the requirement of wind-turbine-specific parameters. The conventional DPC methods are optimum-relation-based MPPT (ORM) paradigms, incremental conductance (INC), and hill-climbing-search-based MPTT (HSM) techniques.
Apart from conventional methods mentioned earlier, a wide range of MPPT techniques have been proposed in the literature; for instance: proportional integral (PI) [67], backstepping controller- [68], fuzzy logic [69], neural network [70], fuzzy-neural network [71], two fuzzy [72], input–output feedback linearization [73], model predictive control [74], adaptive support vector machine combined with firefly algorithm [75], and an adaptive dynamic programming-based MPPT technique [76]. The conventional control techniques for WECS control have traditionally been proportional-integral-derivative (PID) controllers [67]. However, due to their sensitivity to variation in operating points, the nonlinear behavior of WTs, disturbances, and the possibility of faulty scenarios, the conventional control techniques cannot be considered as reliable control techniques. Researchers have proposed various innovative advanced control strategies to improve the WECS control system performance in this regard. Robust control techniques are widely used in the literature to mitigate wind effects and improve power quality, with H 2 and H [77] methods being the most commonly used [78,79]. The MPC also provided an enhanced performance for WECS due to its reduced switching frequency, inherent optimization features, and minimized error. The MPC makes the system mathematically complex and sensitive to parameters uncertainties [80]. Soft-computing-based methods, including artificial neural networks, fuzzy logic control (FLC), and meta-heuristic algorithms-based controllers, are advanced control techniques that provide a quick and efficient response in order to handle the uncertainties in WECS [81,82]. Similarly, the authors of [83,84] created fuzzy-based FTC approaches for WT power control.
WECS is a highly nonlinear system prone to several uncertainties; thus, robust nonlinear control is preferable for smooth operation. Sliding mode control (SMC) is a robust and simple control scheme with a dynamic behavior [85] that was developed in 1977 by Uktin for parametric variations and internal and external unknown disturbance in variable structure systems [86]. SMC applications for WECS start with first-order SMC paradigms, as reported in the literature by [87,88,89,90,91]. First-order SMC schemes are subjected to inherent undesirable phenomena with finite amplitude and frequency called chattering [92]. Conventional SMC is discontinuous, which leads to chattering, singularity, and sensitivity to mismatched uncertainties and disturbances, which restrict its practical application because unmatched uncertainties have an impact on a wide range of practical systems. Furthermore, while asymptotic stability is guaranteed by conventional SMC, there is no assurance that it will occur in finite time, particularly in the presence of unmatched disturbances. When a system is in its sliding mode, it behaves like a predetermined reduced-order system and stops responding to changes in parameters. The researchers presented many methods to mitigate the chattering phenomena and improve the dynamic behavior, resulting in various variants of SMC. Initially, a boundary layer was used by [87] to remove the chattering problem. The many techniques presented are: appropriate finite gains [93], high-order SMC (HOSMC) [94,95], second-order SMC (SOSMC) [96,97,98,99], fuzzy SMC [100], extended state observer-based [101], fractional-order SMC (FOSMC) [102,103], fractional-order terminal SMC (FOTSMC) [104], super twisting FOTSMC (ST-FOTSMC) [105], artificial intelligence integrated FOSMC [106], and composite FOSMC [2]. The SMC variants covered in the literature proposed for WECS are given in Figure 4. Various recent control techniques presented in the literature are compared in Table 1 for various attributes, where ✓ denotes the occurrence of the attribute and ✗ denotes the absence of the attribute.
This paper presents comprehensive insight into the WECS, employing various typologies and SMC strategies to transfer wind power to the electric grid. This paper starts with an introduction of the MPPT techniques for WECS. The comprehensive utilization of various types of generators used in WECS is then discussed and tabularized for comparative analysis. The mathematical modeling of the wind turbine, doubly fed induction generator (DFIG), and permanent magnet synchronous generator (PMSG) is discussed. The general introduction of SMC is discussed and is categorized into (1) integer-order SMC (IOSMC) and (2) FOSMC based on sliding surface construction. In IOSMC, first-order SMC is discussed for WECS, and mathematical proof of the control laws for both the grid side converter (GSC) and rotor side converter (RSC) are presented. The various variants of first-order SMC discussed and mathematically presented are HOSMC, ISMC, BSMC, and fuzzy-SMC. Mathematical study and control law derivation for the three primary variants of FOSMC presented in the literature, including FOTSMC, ST-FOTMSC, and composite FOSMC, are presented.

3. Modeling of WECS and Generators for Closed Loop Control

3.1. Wind Turbine Model

The kinetic energy in wind runs the turbine in order to generate the mechanical power used to drive the generator. This mechanical power is expressed by the following equation as a function of wind speed and wind turbine geometry [105]:
P ω = 1 2 ρ π R 2 C p ( λ , β ) v 3
where
λ = Ω t R v
C p and λ correlate by the expression given as follows:
C p = c 1 c 2 λ 1 e c 3 λ
A maximum value of C p = C p - m a x is achieved when the value of λ equals the optimum value of λ , i.e., λ = λ o p t . The wind turbine torque is given as follows:
T r = T t G Ω t = Ω r G
Putting (4) into (1) and (2), the Ω r - r e f and P g r i d - r e f is given as:
Ω r - r e f = λ o p t G R v P g r i d - r e f = 1 2 η ρ π 2 C p - m a x v 3 }
where η is the wind turbine efficiency.

3.2. Doubly Fed Induction Generator Model

The DFIG has strongly decoupled dynamics; thus, the representation of DFIG in a three-phase system is quite a difficult task. In this section, the DFIG dynamic mode in a dq-reference frame is presented and is given below [105].
V d s = R s I d s + d d t φ d s ω s φ q s V d s = R s I q s + d d t φ q s + ω s φ d s V d r = R r I d s + d d t φ d r ω s ω r φ q s V q r = R r I q s + d d t φ q r ω s ω r φ d s }
where
φ d s = L g I d s + M I d r φ q s = L s I q s + M I q r φ d r = L s I d r + M I d s φ q s = L s I q r + M I q s }
T e m = 3 2 P M V s ω s L s φ q s I d r φ d s I q r J Ω ˙ r = T e m T r f r Ω r }
Here, b = 1 / J , a = f / J , and f = T L J . Aligning the reference frame to the d-axis of the stator flux, we have φ d s = φ s and φ q s = 0 , and, therefore, T e m in (8) can be simplified as:
T e m = p M V s ω s L s φ s I q r
The active power, reactive power, and rotor voltages may be achieved by assuming a constant stator flux, neglecting the per phase stator resistance, taking V d s = 0 and V q r = V s = V d r = ω s φ s , and putting (9) into (6) and (7), and are given as follows:
V d r = R r I d r + σ L r d d t I d r s σ L r ω g I d r V q r = R r I q r + σ L r d d t I q r σ s ω s I d r + s M V s L s }
Here, σ = 1 M L r L s and s = ω s ω r ω s .
P s = M V s L s I q r Q s = V s 2 ω g L s I q r M V s L g I d r }
The d q representation of the rotor current is given as:
d d t I d r = 1 σ L r V d r R r I d r + s σ L r ω s I q r M L s d d t φ d s d d t I q r = 1 σ L r V q r R r I q r s σ L r ω g I d r s ω s M L s d d t φ d s }

3.3. Grid Model

The dynamic model of the grid side is given using the following dynamic equations [36]:
V g d = e d R g I g d L g d d t I g d + ω g L g i g q V g q = e q R g I g q L g d d t I g q ω g L g i g d }
It can be re-arranged as follows:
d d t I g d = 1 L g e d R g I g d V g d + ω g L g I g q d d t I g q = 1 L g e q R g I g q V g q ω g L g I g d }
where the active and reactive power are given as follows:
P g = 3 2 V g d I g d + V g q I g q = 3 2 V g d I g d Q g = 3 2 V g q I g d + V g d I g q = 3 2 V g d I g q }
The current control strategy is derived such that the current x follows the reference current x r e f . The x r e f is given as follows:
x r e f = I d r - r e f I q r - r e f T
The reference q a x i s current is generated using (9), given as:
I q r - r e f = ω s L s P M V s T e m - r e f
The d-axis reference current is derived by substituting Q s - r e f = 0 into reference reactive power Q s - r e f = V s 2 ω s L s M V s ω s L s I d r , and is given as:
I d r - r e f = V s ω s M
The gusty and stochastic wind nature results in a dynamical and variable flow of wind through the DFIG rotor. Thus, an efficient control strategy should be employed to accomplish the non-trivial task of DC link voltage constant regulation. Thus, a vector control strategy was employed here to achieve this task. The vector control strategy aligns the reference frame orientation with stator or grid voltage. Hence, the active and reactive power after V S = V D and V Q = 0 are given as:
P g = 3 2 V g d I d + V g q I g q = 3 2 V g d I d Q g = 3 2 V g q I g d + V d I g q = 3 2 V g d I g q }
I d and I q directly affect the electric power flow between the grid converter and grid as depicted in (19). As the DC power is equivalent to the flow of the active power between the grid side converter and grid, the dynamics can be shown as [106]:
I o s = 3 2 E V g d I g d
C d E d t = I o s I o r
Substituting (21) in (20), we obtain:
E ˙ = 1 c 3 2 E V g d I g d I o r E ˙ = g ( x ) I g d 1 c I o r } where g ( x ) = 1 c 3 2 E V g d
Adding uncertainty term Δ g ( x ) to g ( x ) in (22), the g ( x ) is given as follows:
g ( x ) = g 0 ( x ) + Δ g ( x ) ; g o ( x ) = 1 c 3 2 E r e f V g d
where E r e f is the reference value of E. Putting g ( x ) from (23) into (22) gives us:
E ˙ = g 0 ( x ) I g d 1 c I o r + d E ; d E = Δ g ( x ) I g d
These equations will be further used to model controllers for WECS.

4. Sliding Mode Control Theory

SMC is an accurate, simple, and robust technique used to control and regulate nonlinear systems prone to external perturbations, disturbances, and parametric uncertainties. SMC is designed to operate the system on a surface known as the sliding surface in the desired manner. The proposed SMC-based control law is responsible for keeping the system in the vicinity of the sliding surface. Thus, the basic principle of SMC can be elaborated as a switching function, and control law can converge any point in the system state space to the proposed sliding surface to gradually stabilize the system’s equilibrium point on the design surface [119]. The SMC design consists of two steps: (1) designing a sliding surface and (2) designing a control law using state-space equations. The DFIG is a nonlinear and controlled system represented by the state-space equations as follows:
x ˙ 1 = x 2 x ˙ 2 = f ( x , t ) + g ( x , t ) u }
Here, the system state variables are shown by x 1 and x 2 , and f ( x , t ) and g ( x , t ) represent the nonlinear function and control gain function; the control input is shown here by u. In sliding mode theory, the ultimate goal is the design of an appropriate sliding variable S = S ( x , t ) R so that, during sliding mode S = S ( x , t ) = 0 , a good compensation of system dynamics is achieved. The derivative of S ( x , t ) can be expressed as follows [119].
S ˙ ( x , t ) = φ ( x , t ) + ψ ( x , t ) u
where φ ( x , t ) = S t + S x f ( x , t ) and ψ ( x , t ) = S x ( g ( x , t ) ) . In order to regulate the stability of systems, it is assumed that: A.1 The function ψ ( x , t ) R is uncertain and can be presented as:
ψ ( x , t ) = ψ 0 ( x , t ) + Δ ψ ( x , t )
where ψ 0 ( x , t ) is a positive known function and Δ ψ ( x , t ) is a bounded control perturbation so that d d t ( Δ ψ ( x , t ) ) < ξ , Δ ψ ( x , t ) ψ 0 ( x , t ) Y < 1 , ξ and Y are unknown boundaries. A.2 The function φ ( x , t ) R is presented as follows: φ ( x , t ) = φ 1 ( x , t ) + φ 2 ( x , t ) , with the following bounded conditions: | φ ( x , t ) | ϕ | S | 1 / 2 , φ 1 ( x , t ) δ 1 | S | 1 / 2 , φ ˙ 2 ( x , t ) δ 2 , and the finite boundaries ϕ , δ 1 , δ 2 > 0  [119].

Errors Definition for SMC and Its Variants

The surfaces in SMC are proposed on the basis of the error between the state variable and its reference. Speed, current, and DC link voltages are controllable state components in DFIG-based WECS. The mathematical proofs of SMC and its variants for DFIG-based WECS utilizes three types of errors for both the GSC and RSC side. These errors are: e Ω (the error between the reference generator speed and actual generator speed), e 1 and e 2 (the error between the reference and actual d q currents), and e E (the error between the reference and actual DC link voltage). The mathematical proofs also utilizes the derivative of the error mentioned earlier. The error and their derivatives are given as follows:
For the RSC side, the error between the reference and actual speed is denoted by e Ω and given as:
e Ω = Ω r Ω r - r e f
Taking its derivative and putting the values from (8), we have:
e ˙ Ω = Ω ˙ r Ω ˙ r - r e f e ˙ Ω = T e m J + d 3 Ω ˙ r - r e f }
The RSC control also utilizes the error between the reference and actual d q currents and is given as follows:
e i = x x r e f e i = e 1 e 2 T e i = I d r I d r - r e f I q r I q r - r e f }
The derivative of the above errors after substituting the values from (12) is given as follows:
e ˙ 1 = 1 σ L r V d r R r I d r + s σ L r ω s I q r M L s d d t φ d s I ˙ d r - r e f e ˙ 2 = 1 σ L r V q r R r I q r + s σ L r ω s I d r s ω s M L s d d t φ d s I ˙ q r - r e f }
Hence, (31) can be written as:
e ˙ 1 = G 1 + 1 σ L r V d r 1 σ L r R r I d r e ˙ 2 = G 2 + 1 σ L r V q r 1 σ L r R r I q r } where G 1 = 1 σ L r ( s σ L r ω s I q r M L s d d t φ d s ) I ˙ d r - r e f G 2 = 1 σ L r ( s σ L r ω s I d r s ω s M L s d d t φ d s ) I ˙ q r - r e f } }
The voltage tracking error e E for GSC is given as follows:
e E = E E r e f
The derivative of the DC link voltage error after substituting the values from (24) is given as follows:
e ˙ E = E ˙ E ˙ r e f e ˙ E = g 0 ( x ) I g d 1 C I o r + d E E ˙ r e f }
The error defined in this section will be utilized further to derive the control law of SMC and its variants.

5. Sliding Mode Control for DFIG Based WECS

This section will describe the various techniques and variants of SMC presented in the literature. A detailed literature review portraying the error, surfaces, typology, and remarks are also provided in this section.

5.1. First-Order Sliding Mode Control Schemes

The first-order SMC is the most basic and simple control design for DFIG-based WECS. The SMC provides a suitable compromise between torque oscillations and efficiency of power conversion [119,120,121]. The author in [122] proposed SMC schemes using the first-order SMC, where the surface was chosen as the power tracking error. The reference power is selected as less than the maximum power to keep an energy buffer for frequency change under abrupt load conditions. A test bench developed at NREL FAST is used to validate the proposed scheme. A similar concept was introduced in [118,123] for grid-connected WECS. The first-order SMC schemes mentioned before are of a continuous nature. Utkin in [98] stated that the SMC schemes are inherently affected by an undesirable phenomenon of chattering that proves to be harmful for the system and causes a low control accuracy, high wear and tear of system parts, and high losses. This phenomenon is due to the discontinuous switching control law in the continuous control schemes. The author in [124] suggest an exponential reaching law to eliminate the chattering problem instead of the constant reaching law. The gain in the exponential-reaching-law-based SMC is scheduled according to the magnitude of error. The gain is increased with an increase in error and decreases with a decrease in error. The enhanced reaching law proposed by [124] is given as:
S ˙ = Λ S K D ( S ) | S | γ x sign ( S )
where D ( S ) = α + ( 1 α ) e β x | S | , 0 < α < 1 , and β x > 0 .
It is evident from the above equation that the controller gain is modified between K | S | γ x and K | S | γ x / α according to the magnitude of error. The digital control scheme based on the adaptive reaching law was introduced in [94] to achieve a switch-free control law called digital sliding mode control (DSMC) for DFIG-based WECS. The sliding motions for the digital system were initially investigated by Miloslavjevic in [120]. The authors present the idea of a quasi-sliding mode (QSM) in which the trajectory of the system is along a surface that gives a sliding-like motion instead of a switching surface. The DPC strategy is deployed with a constant switching frequency to improve the power quality and eliminate the need for the synchronous reference frame to minimize the online calculations. The surfaces selected for the design of DSMC are given as follows:
S P = P r e f ( k ) P s ( k )
S Q = Q r e f ( k ) Q s ( k )
This digitalized SMC guarantees sliding at each sampling instant [121,122]. More information on digital SMC can be found in [120,125].
Keeping in view the concept provided by the above-mentioned papers, the control law for the first-order system is derived. The errors and its derivative that are given as follows are used to derive the control law. Taking the surface as the speed error given as follows:
S Ω = e Ω
The derivative is given as follows:
S ˙ Ω = e ˙ Ω = T e m J + d 3 Ω ˙ r - r e f
Now, following the SMC procedure, taking the S ˙ Ω = 0 and solving for equivalent term, one obtains:
T e m e q = J Ω ˙ r - r e f d 3
The RSC speed equivalent control law T e m e q is obtained using (40) and the switching control part T e m s = J k 3 sign ( S Ω ) and is given as follows:
T u = T e m e q + T e m s T e m e q = J Ω ˙ r - r e f d 3 T e m s = J k 3 sign S Ω }
The current control loop of RSC is derived by taking the surface equal to the current errors given as follows:
S = S 1 S 2 = e 1 e 2
The surface derivative is given as:
S ˙ = S ˙ 1 S ˙ 2 = e ˙ 1 e ˙ 2
Now, following the SMC procedure, taking the S ˙ = 0 and solving for the equivalent term, one obtains the following control law for the current loop:
u = h 1 [ R r σ L r I d r s ω s I q r + I ˙ d r - r e f e q u i v a l e n t t e r m k 1 sign S 1 R r σ L r I q r + s ω s I d r + s M V s σ L r L s + I ˙ q r - r e f e q u i v a l e n t t e r m k 2 sign S 2 ]
A similar procedure is followed for GSC. The surface is selected as the difference in the dc link voltage and reference voltage and is given as:
S E = e E
where the error derivative is given as follows:
S ˙ E = e ˙ E
By using the grid side nonlinear model and the procedure followed in the manuscript, the control law can be chosen as:
I d = I d e q + I d s I d e q = 1 g 0 ( x ) E ˙ r e f + 1 C I 0 r I d s = 1 g 0 ( x ) k 4 sign S E }
The operational diagram of the first-order SMC is given in Figure 5. First-order SMC is simple to handle and implement. First-order SMC only requires four gains to be controlled for its operation. Various SMC techniques employed in the literature are tabulated in Table 2.

5.2. Supertwisitng Sliding Mode Control

Leonid Fridman et al. classify the sliding mode control into five generations [127] termed as: (1) first-order SMC, (2) second-order SMC schemes, (3) super twisting algorithms, (4) arbitrary-order SMC, and (5) continuous arbitrary SMC. The super twisting sliding mode control (STSMC) comes under the third generation and has advantages of Lipschitz disturbances with a continuous-type control signal. The authors in [96,97,128,129,130,131] proposed STSMC for WECS. The HOSMC retains the FOSMC merits, including robustness and invariance, with the capability to eliminate chattering and improve the control accuracy, eliminating the relative degree limitations [132]. Various algorithms can be applied to achieve HOSMC, and the most widely accepted is the simplified structured super twisting algorithm (STA), as it requires little target information [133,134]. The following STA structure proposed in [135] is considered here to derive the control law.
U = α S 1 2 sign S + u 1 where u 1 = β sign S }
where α and β are the bounded control gains determined by the uncertain disturbances of the system under investigation. Keeping in view the abovementioned advantages of STSMC, it has proven to be one of the most applied control techniques for the wind-based energy conversion control system. The various systems employing the STSMC scheme with its surfaces, errors, and system types are tabulated in Table 3. The authors in [136] proposed a voltage-mode HOSMC method where a sensorless approach is used to obtain the MPPT optimal voltage reference. The proposed voltage regulator is compared to the conventional SMC voltage controller using two different test cases that use realistic and step change wind speed profiles ranging from 5 m/s to 12 m. The performance of both controllers is presented for the entire operation period at any speed. The HOSM controllers are based on the super twisting algorithm, which requires only the measurement of the sliding variables without the use of information about the sliding constraint’s time derivatives. The quadratic form Lyapunov function method is used to select controller parameters and ensure the closed-loop system’s finite time stability. Simulations under wind fluctuation and load perturbation show that the proposed control strategy is well-suited for controlling the DFIG-WECS and that the control objectives are met in a variety of conditions [130]. The author in [96,131] applies STA for maximum power extraction with finite, reaching time, robustness, and chattering-free control with pre-assumed upper bounds of the externally applied disturbances. The drawback of upper bound pre-assumption was overcome by the authors in [137]. Initially, an adaptive multivariable control scheme was proposed with an adaptive gain adjustment and finite time convergence. A novel Lyapunov stability theory is proposed by [137] to prove the finite time convergence and stability of the proposed control scheme. An adaptive super twisting control scheme was presented in [118] for floating wind turbines. The authors of [131] developed a fault-tolerant SOSMC controller for a DFIG-based 1.5 MW three-blade WT MPPT control subjected to external disturbances and un-modeled dynamics. The authors of [135] designed SOSMC to solve the problem of grid frequency alterations and unbalanced voltage sags. The various systems employing the STSMC scheme with its surfaces, errors, and system types are tabulated in Table 3, whereas the operational diagram is shown in Figure 6.
Using the equivalent term derived in (41) and the STA-based control law, a new HOSMC-based control paradigm is derived as follows:
T u = T e m e q + T e m s T e m e q = J Ω ˙ r - r e f d 3 T e m s = J α ω S Ω 1 2 sign S Ω β ω sign S Ω }
Similarly, for the RSC current control, the equivalent part in (44) is used with the STA-based control law for the derivation of the following RSC control:
u = h 1 [ R r σ L r I d r s ω s I q r + I ˙ d r - r e f + α 1 S Ω 1 2 sign S Ω β 1 sign S Ω R r σ L r I q r + s ω s I d r + s M V s σ L r L s + I ˙ q r - r e f + α 2 S Ω 1 2 sign S Ω β 2 sign S Ω ]
The GSC uses the same surface that comprises the error between the DC link voltage and its reference to transfer the power from the RSC to the grid. Using the STA algorithm proposed in (48) for the switching control part and equivalent law from (47), the new HOSMC law for GSC is given as follows:
I d = 1 g 0 ( x ) E ˙ r e f + 1 C I 0 r λ E S E 1 2 sign S E β E sign S E
The operational diagram of STSMC is given in Figure 6. Keeping in view the above equations, the super twisting SMC increases the complexity of the system. The gains are also increased as the RSC now requires six gains, whereas GSC requires two gains to be tuned. The STSMC-based control schemes improve the system performance in terms of chattering elimination but can reduce the robustness of the system due to the continuous term in the control law.

5.3. Adaptive High-Order Sliding Mode Control

The two controller gains α and β are fixed and chosen to control the performance in STA algorithm-based control laws. In each control law above, these gains are usually determined by the uncertainties boundary, which depends on various factors. The boundary of uncertainty is usually estimated adequately in the control law design process, resulting in unnecessary gains. The issues of unnecessary constant gains are resolved by using adaptive sliding mode control schemes (ASMC). ASMC-based schemes combine SMC theory and adaptive algorithm attributes to adjust the gain to be small enough to maintain the sliding motion. [137] Hao Chen et al. in [119] recently proposed an adaptive STA-based SMC scheme for a PMSG-based tidal stream turbine. The HOSMC scheme mentioned above can be made adaptive using the gains given as follows:
α ˙ = σ γ 1 / 2 , if S 0 0 , if S = 0 β = ε α }
The stability proof for the adaptive STSMC can be found in [119] with the Lyapunov function given by [148], and the operational diagram is shown in Figure 7.
A similar adaptive HOSMC was presented in [137]. A multivariable adaptive multivariable super twisting control scheme for DFIG-WECS MPPT was proposed with unknown upper bounds of disturbances. The proposed scheme has finite-time convergence, chatter-free behavior, and excellent robustness properties against external disturbances and parameter uncertainties. The surfaces are selected as follows:
S 1 = e P ( t ) + K P e P ( τ ) d τ S 2 = e Q ( t ) + K Q e Q ( τ ) d τ }
A new discontinuous control law represented by w was proposed and is given as follows:
w = k 1 S | | S | | 1 2 + r r ˙ = k 2 S | | S | | where w = w P , w Q T and r = r P , r Q T }

5.4. Integral Type Sliding Mode Control

Utkin et al. proposed the ISMC to address the robustness in the reaching phase of conventional SMC [149]. The sliding phase is enforced in the overall system response, whereas the system’s order is maintained in ISMC [106]. Several ISMC-based strategies have been proposed in the literature for WECS to improve the conventional SMC performance tabulated in Table 4. For instance, the author in [150] presented ISMC with a disturbance observer for WECS. The proposed observers estimate the value of the aerodynamic torque as well as the d- and q-axis disturbances. The designed aerodynamic torque observer can estimate the wind speed with fast-changing behavior. The proposed observers’ and controller’s stability is examined using Lyapunov stability theory. The authors in [151] validated ISMC to control the three-blade wind-turbine-based WECS optimally. The analysis of the results reveals that the potential gain in power capture can be achieved by improving the performance of tracking, which always comes at the cost of an increased variation in input torque, which ultimately increases the stress in the system. The main advantage of ISMC is the continuity in the control law in the region 2.5, which is derived for region 2, whereas, for existing baseline controllers, a separate control law is required in region 2.5 for satisfactory operation. A direct power control strategy for WECS based on intelligent ISMC with a sliding mode module is presented in [152] to compensate for the disturbance in the system and circumvent the error in disturbance estimation. The ISMC-based control law comprises both continuous and discontinuous control components to increase the robustness of the system. This controller is made up of two nested controllers: an intelligent proportional integral controller that is supplemented by a sliding mode compensated controller. To derive a typical ISMC scheme for DFIG-WECS, the surface S and its derivative S ˙ presented based on ISMC theory are given as:
S Ω = c 5 e Ω + c 6 e Ω d t S ˙ Ω = c 5 e ˙ Ω + c 6 e Ω }
Following the SMC theory and surface given in (55), the control law comprising an equivalent and discontinuous term is given as:
T u = J Ω ˙ r - r e f d 3 c 6 c 5 e Ω k 3 c 5 sign S Ω
The surface S 1 and S 2 and their derivatives S ˙ 1 and S ˙ 2 for the current control loop, presented on the basis of ISMC theory, are given as:
S 1 S 2 = c 1 e 1 + c 2 e 1 d t c 3 e 2 + c 4 e 2 d t S ˙ 1 S ˙ 2 = c 1 e ˙ 1 + c 2 e 1 c 3 e ˙ 2 + c 4 e 2 }
The resulting control law for the current loop control is now given as:
u = h 1 [ R r σ L r I d r s ω s I q r + I ˙ d r - r e f c 2 c 1 e 1 k 1 c 1 sign S 1 R r σ L r I q r + s ω s I d r + s M V s σ L r L s + I ˙ q r - r e f c 4 c 3 e 1 k 2 c 3 sign S 2 ]
For GSC control, the surface is chosen as follows:
S E = c 7 e E + c 8 e E d t S ˙ E = c 7 e ˙ E + c 8 e E }
By using the grid side nonlinear model and the procedure for SMC law derivations, the control law can be given as:
I d = 1 g 0 ( x ) E ˙ r e f + 1 C I 0 r c 8 c 7 e E k 4 c 7 sign S E
Based on the above derivations, the overall control scheme for RSC needs nine gains to be tuned, whereas three gains are needed for the GSC side control. The operational diagram of ISMC is given in Figure 8.

5.5. Terminal Sliding Mode Control

Well-documented research works highlights the SMC theory and application, such as [125,162,163]. The asymptotic stability of SMC structures is based on Lyapunov theory underpinned by the Lipschitz condition for ordinary differential equations (ODEs). The asymptotic stability nature shows a slow state convergence rate of the system dynamics near the equilibrium [164]. This depicts that the finite-time convergence to the equilibrium point is not possible in finite time. The author in [165] achieved finite-time convergence and utilized a terminal attractor as a first-order dynamics with fractional power to add some non-smoothness in the neural learning rule. The authors in [166] extended the terminal attractor idea to SMC schemes and proposed a terminal SMC (TSMC) scheme for the second-order system. Similarly, the authors in [167] proposed TSMC for a single-input–single-output (SISO) of higher order and the same applies for authors in [168,169] for multi-input–multi-output systems. The singularity problems in TSMC schemes were solved in the early 2000s [170], followed by fast TSMC [171] and continuous TSMC [172]. In the previous 5 years, the TSMC enjoyed an over 20% annual growth in the theoretical and application contributions [164]. The authors in [166] used the following form of terminal attractors to control the second-order system:
S = x ˙ + β | x | λ s i g n ( x )
where x is a variable and, β > 0 and 0 < λ < 1 . Initially, the value of λ was taken as q/p, where q and p are positive integers. The time for x to reach 0 when x ( 0 ) 0 and s = 0 is given as:
t s = β 1 ( 1 λ ) 1 | x ( 0 ) | 1 λ
The author in [171] proposed a fast TSMC (FTSMC) and increased the convergence rate by modifying (61) as follows:
S = x ˙ + α x + β | x | λ s i g n ( x )
where the convergence time is given as:
t s = α 1 ( 1 λ ) 1 l n α | x ( 0 ) | ( 1 λ ) + β l n ( β )
During the sliding phase, when S = 0, the Equation (64) becomes as follows:
x ˙ = α x β | x | λ s i g n ( x )
Equations (64) and (65) shows that, for FTSMC, the system will reach x = 0 faster than (61). The SMC design process used derivatives of the sliding surface, which can cause the problem of singularity due to the presence of λ 1 in the time derivative of β | x | λ s i g n ( x ) , i.e., β λ | x | λ 1 s i g n ( x ) x ˙ , leading to infinite control. A nonsingular TSMC proposed by [170] overcame the problem of singularity given as follows:
S = β x + | x ˙ | λ s i g n ( x ˙ )
Here, 1 < λ < 2 . For the system having a relative degree >1, the singularity problem is overcome by a modified version of TSMC known as integral TSMC (ITSMC) and is given as follows:
S ( t ) = x ( t ) + β 0 t | x ( τ ) | λ s i g n ( x ( τ ) ) d τ
Keeping in view the above theoretical considerations of TSMC, a major contribution has been added in the literature for the application of TSMC in DFIG-based WECS. A thorough literature review of TSMC for DFIG-based WECS is tabulated in Table 5. For instance, the author in [173] proposed adaptive FTSMC for DFIG-based WECS to mitigate the chattering in the system dynamics. The author in [174] proposed ITSMC for the direct power control of DFIG-based WECS. A discrete time TSMC is proposed in [175] for PMSG-based WECS, where the authors achieved an efficiency of 94 % . The authors in [173] applied a non-linear FTSMC to obtain faster convergence. The authors developed a higher-order SMC to address the control problem of DFIG-based WECS in a variety of challenging situations. To reduce chattering, they used a fast adaptive TSMC with a nonlinear sliding surface. The proposed controller demonstrated a greater robustness to wind speed variations and external disturbances than the feedback linearization and conventional PI methods. The surface utilized is given as follows:
S i = β i 1 e i + β i 2 | e i ˙ | λ 1 s i g n ( e i ˙ ) + β i 3 | e i ˙ | λ 2 s i g n ( e i ˙ )
Authors in [174] proposed a strategy that used the abc frame of reference, and its gain parameters were dynamically defined based on the absolute value of the respective sliding surfaces. The analysis of the results showed that the proposed controller for the DFIG is very effective in reducing inter-area oscillations in the multimachine network at both low and high wind speeds. The robustness and stability of the proposed control strategy for both the RSC and the GSC are validated by a time-domain mathematical simulation in the MATLAB/EDITOR environment, and a practical simulation in the MATLAB/SIMULINK environment. The test cases selected included “sub” and “super” synchronous mode of operation of the DFIG and “lower” and “higher” wind speed operations for the stator reactive power reference and random wind speed variation. In order to reduce chattering, the authors proposed a nonlinear terminal sliding surface, which is given as follows:
S ( t ) = e ( t ) + β 0 t | x ( τ ) | λ d τ
The authors in [176] investigated the three-phase GSC control problem of variable speed PMSG-based WECS. To control the active and reactive powers exchanged between the converter and the grid, the authors proposed two integral-type TSMC controllers with attenuated chattering. They used integrators to soften the switching signals and generate continuous control signals to reduce chattering. The controller’s time-varying gain is designed to reduce extra control effort and avoid overestimation of system uncertainties. Virtual control is used to regulate the GSC’s input currents by forcing the voltages in the GSC’s three capacitors to track the virtual signals. The controller design takes into account both matched and unmatched parametric uncertainties to ensure the system’s robustness.

5.6. Fuzzy logic Sliding Mode Control

Fuzzy logic control (FLC) has added many features to express human thinking ambiguities by supplanting conventional techniques. FLC has a good feature of dealing with an unknown method [178]. The estimation of system parameter uncertainties increases the adjustability of SMC’s switching gain, resulting in chattering attenuation and controller robustness [179]. FLC theory has been employed to SMC in several ways. Certain problems including chattering in the sliding phase, state estimation, discontinuous part estimation, signum function approximation, and complex have tuning in SMC and its variants have been mitigated using FLC theory. For instance, ref. [180] proposed a FLC scheme for a PMSG based WECS. The authors used a boost converter to control the dc link voltage. The input power and dc link current is used to approximate the dc link voltage. This FLC’s basic operational principle is to perturb the DC-side voltage Vdc and observe the corresponding variation in the DC-side power Pdc. If the power increases with the most recent voltage increase, the search will continue in the same direction. If, on the other hand, the voltage increase reduces the DC-side power, the searching process will be reversed. The proposed FLC-based system is shown in Figure 9. The same scheme is proposed by the authors in [181], but, here, the input power and voltage were used to calculate the dc link current using FLC theory. The authors in [182] utilized the FLC theory and approximated the discontinuous part of the SMC scheme and improved the SMC performance by replacing the chattering source by FLC. The drawbacks of robust controllers were avoided by using this design, where, to eliminate chattering, the fuzzy-based variable switching gain technique was introduced. The same approach is used by [183] and the signum function is replaced by FLC to mitigate the chattering phenomenon. Moreover, the membership function is adjusted using the PSO-GSA optimization algorithm. Using the scheme presented in [183], an FLC theory applied to SMC is explained here.
The authors selected the tracking error e j as an input to the fuzzy system v j , h = ρ j sign e j as its output. In the next step, the fuzzy rules are decided and are given as follows:
Rule 1 : If S j is P, then u j , h is PE ;
Rule 2 : If S j is Z, then u j , h is Z E ;
Rule 3 : If S j is N, then u j , h is NE ;
where the output variable v j , h fuzzy labels are Z: Zero, N: Negative, P: Positive and the fuzzy labels for the input variables S j are ZE: Zero effort, NE: Negative effort, and PE: Positive effort.
The Max-Min method is used to perform the inference and thus the control law can be expressed as follows:
v j , h = i = 1 3 μ j , i ρ j , i i = 1 3 μ j , i = μ j , 1 ρ j , 1 + μ j , 2 ρ j , 2 + μ j , 3 ρ j , 3
where 0 μ i , 1 1 for i = 1 , 2 , 3 are the firing strengths of the three fuzzy rules; and ρ j , 1 = ρ j , ρ j , 2 = 0 and ρ j , 3 = ρ j are the center of the output membership functions ZE, NE, and PE, respectively. Various cases will be triggered depending on the input to the fuzzy system. Thus, the fuzzy-based control law can be expressed as follows:
T u = T e m e q + T e m s T e m e q = J Ω ˙ r - r e f d 3 T e m s = J v j , h }
For the RSC current control, the control law will take the following form:
u = h 1 [ R r σ L r I d r s ω s I q r + I ˙ d r - r e f v j , h R r σ L r I q r + s ω s I d r + s M V s σ L r L s + I ˙ q r - r e f v j , h ]
By using the grid side nonlinear model and the procedure followed in Section 5.1, the control law under a fuzzy control structure is given as:
I d = I d e q + I d s I d e q = 1 g 0 ( x ) E ˙ r e f + 1 C I 0 r I d s = 1 g 0 ( x ) v j , h }
Without requiring a detailed mathematical model, a fuzzy logic controller can track the maximum power point more efficiently, especially in the presence of frequently changing wind conditions. This heuristic-reasoning-based automatic control approach has the added benefits of adaptability, simplicity, extending operating ranges, and being insensitive to parametric uncertainties [180]. The various fuzzy-based control schemes are tabulated in Table 6.

6. Passivity and AI Based Sliding Mode Control

The passivity-based controllers gained attention due to the energy-based approach used to guarantee stability. It reshapes the natural energy of the system by considering the energy properties and injecting the damping terms. An SMC strategy based on the integral control theory and considering the energy properties of brush-less DFIG was proposed in [173] to regulate the power factor and active power. A Hamiltonian model of DFIG was developed on the basis of a seventh-order model and then used to propose a robust SMC-based control scheme. The surface selected
s e x = H v ( e x ) + σ ( e x )
σ ( e x ) is the integral variables vector.
H v is proposed on the basis of the Hamiltonian function given as:
H = H m + H e H m = 1 2 J m J m Ω m Ω m 2 H e = 1 2 φ s φ s T A φ s φ s φ s }
The performance of the proposed control scheme was validated by comparing with PI and the passive controller and by using four types of perturbations, which include parameter, mechanical torque, interconnection voltage, and short circuit variations. The power factor stabilizes after 2 s of perturbation with no overshoot. The passivity approach was also implemented in [174] for PMSG-based WECS. The authors construct the storage function to achieve MPPT, which is the combination of the resistance produced by the resistor heat produced by the d-axis current, shaft system kinetic energy, and torque energy. The energy function constructed is given as follows:
H i d , Ω m , T e , T m 1 2 i d i d 2 resistor heat + 1 2 Ω m Ω m 2 kinetic energy + 1 2 T m T e J tot Ω m 2 accelerating torque energy
The derivative of this function is taken and, considering SMC theory, the passivity-based SMC is designed.
To further increase the robustness of the system, the authors in [184] propose an artificial neural network (ANN)-based SMC in which a torque compensation term is fed forward. The input to the ANN network is chosen to be the error between the actual and optimum speed. The SMC with ANN has a lower speed error of 0.67 rad/s and lower maximum power tracking error of 120 W compared to the PI controller with a speed error of 1.2 rad/s and max power tracking error of 145 W. The passivity and AI-based SMC schemes reviewed in the literature are given in Table 7. The overall SMC-based passive control system improves the system robustness and stability under stochastic wind speed variations, varying pitch angles, and generator parametric uncertainties.

7. Fractional-Order Sliding Mode Control Schemes

Fractional-order control theory has vast emerging applications in engineering, biosciences, and medical sciences [185,186], with a greater degree of freedom for integer-order systems [187]. SMC has excellent performance characteristics, such as a fast response, high robustness against external disturbances, and computational simplicity, making it a viable solution for a wide range of linear and nonlinear systems [77,188,189,190,191]. Because conventional design methodologies consider linear sliding surfaces, finite-time convergence of system states to equilibrium may fail [189]. As a result, the reaching gains should be chosen to be large enough when the SMC surface is close to equilibrium. In order to fill this void, TSMC approaches have been developed [192,193]. To ensure the system dynamics’ finite-time convergence, TSMCs incorporate nonlinear functions into the SMC surface. Although TSMC has a faster finite-time convergence and higher control precision than conventional SMC, it has two major drawbacks [194]: (a) the singularity problem, which causes unbounded large control inputs and has been well addressed by nonsingular TSMC (NTSMC) modifications [195,196], and (b) the chattering phenomenon, which is caused by high-frequency control switching. Fractional-order (FO) calculus has recently demonstrated a promising performance in improving conventional control approaches and has been widely applied in many research areas [197,198,199,200]. To address the second disadvantage, various combinations of SMC methods and FO calculus have been proposed in the literature [201,202,203]. When compared to traditional SMC and TSMC approaches, the integration of FO calculus with SMC provides more degrees of freedom by introducing more design parameters into the system. The FO calculus adds a memory to the controller [198], allowing it to consider the entire history of input signals, effectively reducing chattering and tracking errors in conventional SMC [203,204]. The authors in [186,205] give insight into the fractional-order integrals and differentiation. A stability analysis was performed in [206], whereas the fractional-order Lyapunov theorem was derived for fractional-order controllers in [207,208,209,210,211,212,213,214]. The primary fractional calculus operator is defined as follows:
a D t α d α d t α , R ( α ) > 0 1 , R ( α ) = 0 α t ( d τ ) α , R ( α ) < 0
A primary fractional operator has three definitions, namely, α th-order Riemann–Liouville fractional, α th-order Caputo fractional derivative, and α th-order Grunwald–Letnikov, and are given as follows:
a D t α f ( t ) = d α d t α f ( t ) = 1 d m Γ ( m α ) d t m α t f ( τ ) ( t τ ) 1 α d τ a D t α f ( t ) = I α f ( t ) = 1 Γ ( α ) α t f ( τ ) ( t τ ) 1 α d τ }
a D t α D α = 1 d m Γ ( n α ) α t f n ( τ ) ( t τ ) α n + 1 d τ ( n 1 α < n ) d m d t n f ( t ) ( α = n )
a G L D t α f ( t ) = lim h 0 1 h α j = 0 [ ( t α ) / h ] ( 1 ) j α j f ( t j h ) α j = Γ ( α + 1 ) Γ ( j + 1 ) Γ ( α j + 1 ) }
where m is the smallest integer number and its value is greater than α . The given subsections will provide mathematical insight into the various FOSMC schemes presented in the literature. The main schemes evident in the literature and presented here are FOSMC schemes, FOTSMC, FO-STSMC, and fuzzy-FOSMC schemes.
Table 7. Passivity-based SMC comparison in WECS.
Table 7. Passivity-based SMC comparison in WECS.
TechniqueRef.YearErrorsSurfacesGenerator, Control, and Converter-Type HardwareRemarks
Passivity-based SMC [215]2021 e x = x x s s where x = [ J Ω r ϕ s ] , x s s = stationary value of x s e x = H v e x + σ e x , where s e x = s 1 e x s 2 e x T , H v e x = H e x H e x T and σ e x = σ 1 e x σ 2 e x T , the H v e x is the energy function obtained using energy-based model of DFIG.Brushless DFIG
RSC control
No
An energy or passivity-based SMC is presented in this paper. Energy-based model of brushless DFIG is also presented, and practical implementation issues are also depicted in the article.
    [216]    2018 e x = x x s s
where x = [ i d , ω m , T e , T m ]
S i d , ω m , T e , T m = 1 2 i d i d 2 resistor heat + 1 2 ω m ω m 2 kinetic energy + 1 2 T m T e J tot ω ˙ m accelerating torque energy Multipole
SG
Yes
A passivity-based SMC is presented in this paper. Extensive analysis was conducted in this paper to evaluate the passivity-based SMC performance.
AI-based SMC  [184]  2014   e Ω = Ω r e f Ω r S Ω = Ω r e f Ω r IG
RSC control
No
Feed forward ANN is used to estimate the sigmoid function in the discontinuous control part of SMC.
 [217]2018 e i = X r d q I r d q - r e f S i = X r d q I r d q - r e f where X r d q are the states estimated using high-order neural network trained by extended Kalman filterDFIG
RSC control
No
A neural network of high order is used and trained online using extended Kalman filter. The estimated states are then used to proposed a SMC scheme.

7.1. Fractional-Order Sliding Mode Control Design

In this control scheme, the reference signals have been generated using the control law driven by fractional-order surfaces. The authors in [102] utilize the error between the reference speed and actual speed to generate the reference torque. The electromagnetic torque in the outer loop generated by FOSMC is compared with the PI control scheme and shows a superior performance in terms of tracking under uncertainties. Similarly, the d q current error is used to propose a sliding surface. The sliding surface proposed by [107] is given as:
S = e + c D α e ;
The proposed control system is made adaptive by estimating the lumped uncertainty using a low pass filter with unity steady state gain. The control law consists of an equivalent part with non-linear known terms, D 1 α e , and a robust term. Similarly, the authors in [218] use the same fractional-order control theory but use direct power control for speed and power tracking. The authors use varying parameters to choose the fractional-order for the precise tracking and chattering level. The disturbance comprising 10% of fifth-order and 8 % of seventh-order harmonics is efficiently tackled with less chattering compared to integer-order SMC. The active and reactive power errors are used to propose the surface given as S = c D α e ; , which generates the required voltage signals for the PWM block.
The same idea has been extended in order to solve the sub-synchronous oscillation phenomenon caused by the DFIG system in multi terminal DC systems in [219]. The switching surface is selected on the basis of the d q current error. The sub-synchronous damping controller is also employed with the FOSMC current controller, which comprises a filter, amplitude gain segment, phase compensation, and proportion link. Thus, the overall control scheme provides a faster convergence rate of 0.4 s at 7.5 m/s under sub-synchronous oscillation variation compared to a conventional damper. Extending the idea of FOSMC, the authors in [103] proposed a FOSMC scheme for enhancing the output power quality. A surface is selected as:
S = e + λ D α 1 s i g n ( e ) γ
The controller coefficent selection problem is solved here by using the Luapunov theorem, where the optimization is used to select optimum parameters. The direct current control typology is employed, where the reference current is generated using an MPPT strategy. The FOSMC shows a superior performance compared to SMC and PI control schemes using simulation and experimental results. Moreover, the FOSMC shows fewer harmonics and a faster steady state convergence with chattering elimination and parameter tuning in [218]. The generic operational diagram of FOSMC controller is given in Figure 10, and the techniques reviewed are presented in Table 8, whereas the mathematical derivation of the FOSMC-based current control scheme is given below.
A surface is proposed using the fractional-order calculus presented as:
S Ω = e Ω + c 3 D α e Ω S ˙ Ω = e ˙ Ω + c 3 D 1 α e Ω }
Equation (83) can be further simplified by using the D α operator on both sides and then putting the value of e ˙ Ω as follows:
S ˙ Ω = T e m J + d 3 Ω ˙ r - r e f + c 3 D 1 α e Ω
Equation (84) is used to derive the equivalent control law T e m e q , and the T e m s (switching control part) is designed using the SM theory given below as:
T u = J d 3 Ω ˙ r - r e f + c 5 D 1 α k 3 s i g n ( S Ω )
A new sliding surface presented using the fractional-order calculus is presented as:
S 1 = e 1 + c 1 D α e 1 S ˙ 1 = e ˙ 1 + c 1 D 1 α e 1 S 2 = e 2 + c 2 D α e 2 S ˙ 2 = e ˙ 2 + c 2 D 1 α e 2 }
Using the surface derivative given in (86) and following the SMC control law methodology, the RSC current control is obtained as follows:
u = h 1 [ R r σ L r I d r s ω s I q r + I ˙ d r - r e f c 1 D 1 α e 1 k 1 sign S 1 R r σ L r I q r + s ω s I d r + s M V s σ L r L s + I ˙ q r - r e f c 3 D 1 α e 2 k 2 sign S 2 ]
By defining a fractional-order surface S E and its derivative S ˙ E , we have:
S E = c 7 D α e E + c 8 D α e E γ sign e E
S ˙ E = c 7 D 1 α e E + c 8 γ D α e E γ 1 e ˙ E
Applying fractional operator D α to (85), we have:
D 1 α S E = c 7 D 1 2 α e E + c 8 γ e E γ 1 e ˙ E
Putting the value of e ˙ E from (34) in (90), we obtain:
D α ¯ S E = c 7 D 1 2 α e E + c 8 γ ( | e E | γ 1 g 0 ( x ) I d | e E | γ 1 C I o r + | e E | γ 1 d E | e E | γ 1 E ˙ r e f where D 1 α = D α ¯ }
Equation (91) is used to derive the equivalent control law I d e q , while the discontinuous control component I d s is designed based on fuzzy control theory and is given as follows:
I d e q = 1 g 0 ( x ) | e E | 1 γ | e E | γ 1 C I o r | e E | γ 1 E ˙ r e f c 7 c 8 γ D 1 2 α e E I d s = λ E | e E | 1 γ g 0 ( x ) c 8 γ k 4 sign ( S 2 ) }
Thus, the complete control law, I d = I d e q + I d s , is derived using (92). The FOSMC basic operational diagram is shown in Figure 10.

7.2. Fractional-Order Terminal Sliding Mode Control Schemes

The advantages of FOSMC have been presented in the previous section. The finite time convergence capabilities of TSMC have been combined with FOSCM to achieve a fast, finite time convergence, and an enhanced performance with minimized chattering. The authors in [104] proposed an adaptive FOTSMC scheme for DFIG-based WECS. The new surface adapted is given below as:
S = c 5 D α e + c 6 D α | e | γ s i g n ( e ) .
The new surface now follows the characteristics of FOSMC and TSMC schemes. The controller has also been made adaptive in order to estimate the disturbance and external perturbations, and the parameters have been selected using the optimization toolbox and integral absolute error as the performance index. The comparison of the results with SMC and PI showed that the FOTSMC showed a robust performance under normal and perturbed situations, with less chattering. Following the same pattern, the authors in [106] integrated the AI with an FOTSMC control scheme to propose a new controller for a more robust performance. The discontinuous parts of the controllers in [109] may increase the chattering due to the presence of D 1 α ; thus, it has been estimated and tuned by FLC theory and the performance has been validated by using PIL results. The reactive power oscillations were ± 0.01 k V A R for SMC, whereas AI-based FOTSMC gives almost zero oscillations. The FOTSMC operational diagram is shown in Figure 11, whereas the techniques have been reviewed in Table 9. The further section will now present the mathematical formulation of FOTSMC proposed by [104]. Following the surface in (93), the derivative is given as:
S ˙ Ω = c 5 D 1 α e Ω + c 6 D α D 1 e Ω γ s g n e Ω S Ω ˙ = c 5 D 1 α e Ω + c 6 γ D α e Ω γ 1 e Ω }
Equation (94) can be further simplified by using the D α operator on both sides, and then putting the value of e ˙ Ω as follows:
D α ¯ S Ω = c 5 D 2 2 α e Ω + c 6 γ | e Ω | γ 1 T e m J + | e ω | γ 1 d 3 | e Ω | γ 1 Ω ˙ r - r e f | where D α ¯ = D 1 α }
The equivalent control law T e m e q is derived from (95), whereas the T e m s (switching control part) is given below as:
T u = T e m e q + T e m s w T e m e q = J | e Ω | 1 γ | e Ω | γ 1 Ω ˙ r - r e f c 6 γ c 5 c 6 γ D 1 2 α e Ω T e m s = J | e Ω | 1 γ λ ω c 6 γ sign ( S Ω ) }
The sliding surface for the current control loop using the fractional-order calculus is presented as:
S 1 = c 1 D α e 1 + c 2 D α e 1 γ sign e 1 S 2 = c 3 D α e 2 + c 4 D α e 2 γ sign e 2 }
The equivalent control laws [ V d e q V q e q ] are derived using the same process for the speed control, and the switching control components [ V d s V q s ] are designed based on a SMC theory and are given below as:
V d e q = σ L r [ G 1 + 1 σ L r R r I d r c 1 | e 1 | 1 γ D 1 2 α c 2 γ e 1 ] V d s = λ d | e 1 | 1 γ c 2 γ s i g n ( S 1 ) V q e q = σ L r [ G 2 + 1 σ L r R r I q r c 3 | e 2 | 1 γ D 1 2 α c 4 γ e 2 ] V d s = λ q | e 2 | 1 γ c 4 γ s i g n ( S 2 ) }
The equivalent terms [ V d e q V q e q ] and switching terms [ V d s V q s ] combine to formulate the complete current control law u. By defining a fractional-order surface S E and its derivative S ˙ E , we have:
S E = c 7 D α e E + c 8 D α e E γ sign e E
Similarly, the control law for grid side is given below as:
I d e q = 1 g 0 ( x ) | e E | 1 γ [ | e E | γ 1 C I o r | e E | γ 1 E ˙ r e f c 7 c 8 γ D 1 2 α e E ] I d f = λ E | e E | 1 γ g 0 ( x ) c 8 γ sign ( S E ) }
Thus, the complete control law I d = I d e q + I d f is derived using (100). The operating diagram of FOTSMC is shown in Figure 11.

7.3. Fractional-Order Super Twisting Sliding Mode Control Schemes

The FOSMC eliminates the chattering by minimizing the decaying time during the sliding motion. This enhanced performance is achieved by implementing fractional-order calculus applications during the sliding motion. However, it has been mentioned in [105] that the chattering in FOSMC can be further minimized by using the HOSMC scheme in the discontinuous control part. The authors in [105] combine the attributes of the super twisitng control structure and fractional-order control theory. The author improves the FTSMC performance by using the equivalent terms in (89), (93), (91), and the super-twisting-based switching control law given in Section 5.2. The operational diagram of FOTSMC is shown in Figure 12 and the reviewed techniques are presented in Table 9. The speed control is given as follows:
T u = T e m e q + T e m s w T e m e q = J e Ω 1 γ e Ω γ 1 Ω ^ r - r e f c 6 γ c 5 c 6 γ D 1 2 a e Ω T e m s w = J e Ω 1 γ λ ω 3 c 6 γ S Ω 1 2 sign S Ω β ω 3 e Ω 1 γ c 4 γ sign S 2 }
The current control law for ST-FTSMC is given as follows:
V d r e q = σ L r G 1 + 1 σ L r R r I d r c 1 e 1 1 γ D 1 2 α e 1 c 2 γ V d r s w = λ ω 1 e 1 1 γ c 2 γ S 1 1 2 sign S 1 β ω 1 e 1 1 γ c 2 γ sign S 1 V q r e q = σ L r G 2 + 1 σ L r R r I q r c 3 e 2 1 γ D 1 2 a e 2 c 4 γ V d r s w = λ ω 2 e 2 1 γ c 4 γ S 2 1 2 sign S 2 β ω 2 e 2 1 γ c 4 γ sign S 2 }
For GSC control, the ST-FTSMC-based control law takes the following structure:
I d e q = 1 g 0 ( x ) e E 1 γ e E γ 1 C I o r e E γ 1 E ˙ r e f c 7 c 8 γ D 1 2 α e E I d s w = λ ω 4 e E 1 γ g 0 ( x ) c 8 γ S E 1 2 sign S E B ω 4 e E 1 γ g 0 ( x ) c 8 γ sign S E }
Thus, the complete control law I d = I d e q + I d f is derived using (103). The operating diagram of FOTSMC is shown in Figure 12.

7.4. Fractional-Order Fuzzy Sliding Mode Control

The signum function in (96), (98), and (100) is the main cause of the chattering in the FOTSMC paradigm. The authors in [2] replace the sign ( S ω ) by the saturation function to minimize the chattering; this is given as:
sat S Ω ψ = S Ω ψ , if | S Ω | | ψ | sign ( S Ω ) , othewise
where ψ is defined as the thickness of the boundary layer. The utilization of the above function in place of sign ( S ) can eliminate the chattering, but the performance may be degraded and not suitable under uncertain scenarios. Therefore, the function sat is replaced by a fuzzy system, which performs like a sat function with a nonlinear slope inside the thin boundary layer. The inputs to the fuzzy system are selected as the fractional-order surfaces S = ( S Ω , S 1 , S 2 , S E ) and alteration in the surface S ˙ = ( S ˙ Ω , S ˙ d , S ˙ q , S ˙ E ) adopted in this paper for the DFIG control. The fuzzy system for the speed control loop can be represented as follows:
u f Ω = F u z z y F O S M C ( S Ω , S ˙ Ω )
where fuzzy−FOSMC is the system developed on the basis of fuzzy control theory using the membership function given in Figure 13. Hence, a new fuzzy-based reaching law can be presented as:
T e m f = J | e Ω | 1 γ λ ω c 6 γ u f Ω
The u f q and u f q are the approximated signum functions using fuzzy theory. The rules map is shown in Figure 13a. Membership functions presented in Figure 13b,c are used to calculate u f q and u f q and are given as follows:
u f d = F u z z y F O S M C ( S 1 , S ˙ 1 ) u f q = F u z z y F O S M C ( S 2 , S ˙ 2 ) }
The u f E is the approximated signum functions using fuzzy theory. Membership functions presented in Figure 13 are used to calculate u f E and are given as follows:
u f E = F u z z y F O S M C ( S E , S ˙ E )
Hence, using the above given structures, the control law given in Section 7.2 receives the attributes of fuzzy control theory to remove the chattering in the sliding phase. The same concept can also be applied for other variants of SMC to improve the performance.

8. Comparative Analysis of Integer and Fractional-Order SMC Schemes

The various typologies are made operational using various control levels. The optimal operation of grid-connected typologies is performed in four levels of centralized control, which are: (1) power converter control, (2) generator and grid control, (3) grid integration and MPPT control, and (4) wind turbine centralized control. Level 4 consists of both electrical and mechanical control. The mechanical control includes damping control, whereas the electrical control includes the reactive power generation (RPG) and fault ride-through (FRT). The tower mechanical resonance in the drive train, electrical resonance in the grid, and torsional vibrations in the drive train are minimized by the electrically and mechanically managed damping control part. Ancillary services, such as spinning reserve, kinetic energy storage, and inertia emulation, are also provided in control level 4 using electrical and mechanical control. The MPPT and grid synchronization is performed in control loop 3. The phase-locked loop (PLL), zero-crossing detection, and grid voltage filtering are used to synchronize the WECS with the grid. The outputs and inputs to the control level 3 are given in Figure 14, where the reference dc-link voltage or grid voltage magnitude is kept constant. The SMC and its enhanced techniques are implemented in level 2 and level 1 using the various mathematical control laws derived in this paper. The FOSMC and IOSMC schemes are compared in the literature using mathematical and numerical analysis. This section will initially provide a mathematical comparison of IOSMC and FOSMC schemes and SMC variants proposed in the literature. This section will also compare the FOSMC and IOSMC schemes for DFIG speed, power, DC link voltage, errors, and sliding surfaces.

8.1. Mathematical Comparison

This section presents the FOTSMC scheme advantages over integer-order SMC (IOSMC) schemes in terms of stability. The IOSMC and FOTSMC system satisfies the following conditions at the occurrence of sliding motion [220]:
D t α f ( t ) = G x ( t ) ( 0 < α 1 ) , G R
The above equation solution can be given as follows:
f ( t ) = X α , 1 ( t ) f 0
where X α , β ( t ) = k = 0 G k t α k / ( Γ ( α k + β ) ) ( 0 < α 1 ) is the states trans- fer function, and Γ ( z ) is the gamma function. For an integer-order system with r = 1 , the state transfer function is given as follows:
E α , β ( t ) = X 1 , 1 ( t ) = k = 0 G k t k Γ ( k + 1 ) = k = 0 G k t k k ! = exp ( G t )
The above result shows that the decay type of integer-order system towards 0 is exponential. For a fractional-order system, the state transfer function from Matignon is given as [220]:
X α , 1 ( t ) = k = 0 G k t α k Γ ( r k + 1 ) 1 Γ ( 1 α ) ( G ) 1 t α ( 0 < α < 1 )
The fractional-order system decays like t α towards 0, depicting the slow energy transfer with FOTSMC surfaces. The delay time Δ defined as the time taken for control output to vary from + u ( x , y ) to u ( x , y ) is the major reason for chattering. A system reaching from initial state ( x 0 , y 0 ) to sliding surface S ( x , y ) = 0 is shown in Figure 15. According to the decaying time proof in Equation (111), the Figure 15 shows the behavior of the integer-order system decaying with e t taking a total time of t i , as shown by the red line. On the other side, the fractional-order system decaying time is t α , as proven in (112), and takes a total time of t f , as shown through the blue line. Thus, for the same plant, e t i > t f α [220].

8.2. Case Study: Implementation and Numerical Comparison

This section presents a comparative analysis of the IOSMC, FOTSMC, and composite FOSMC control schemes presented in the literature. The detailed operational diagram of FOTSMC presented in Section 7.2 is shown in Figure 16. The authors in [2,105] provided several SMC schemes and their variants. The resulting speed waveform and active and reactive power for SMC, FOTSMC, and fuzzy FOSCMC are compared in Figure 17. Figure 17a shows that the controllers provided an accurate speed tracking, with SMC providing sever chattering. An FOTSMC-based controller was developed to minimize the chattering in the conventional SMC. The results depicts that the FOTSMC minimized the chattering in the SMC scheme to a certain extent. To further improve the SMC and FOTSMC performance, the author in [2] proposed a fuzzy-FOSMC scheme. The results in Figure 17 show that the proposed fuzzy-FOSMC further improved the performance of DFIG-WECS by minimizing the chattering effect in SMC and FOTSMC. The error and surfaces waveforms in Figure 18 also show that the fuzzy-FOSMC provided an improved performance compared to the SMC and FOTSMC. A computational comparison was evaluated in this paper using PIL-based simulation, which consists of a dual core processor TMS320F379D programmed using a rapid prototyping method from a Simulink environment. The Simulink files are compiled in the discrete domain and the hex file is programmed into the random access memory (RAM) of the processor. It can be seen from Figure 19 that the fractional SMC has more CPU utilization compared to the integer-order SMC. The average CPU utilization for fraction SMC is 29.13%, whereas, for SMC, the average CPU utilization is 18.81%. Similarly, the average execution time of fractional SMC is also higher than the integer-order SMC as shown in Figure 19.

9. Conclusions

A comprehensive review of sliding mode control and its variants presented in the literature is presented in this paper. The main focus of this paper is to cover the various SMC techniques proposed and mathematically presented in the literature. The important factor in a wind energy conversion system is the mathematical modeling of various components. The mathematical model of the wind turbine, doubly fed induction generators, permanent magnet synchronous generators, and grid is presented. This study starts with the WECS overview and the control levels in a wind-energy-based grid system. The rotor side control (RSC) and grid side control (GSC) concept is used to assess the mathematical and operational aspects of grid-connected WECS. The SMC techniques are broadly classified into integer-order SMC (IOSMC) and fractional-order SMC (FOSMC). The IOSMC is then classified based on the order and combination with other techniques, which include terminal SMC, high-order SMC, and composite SMC. The same classification is then applied to FOSMC variants. The mathematical proof, including the proposed errors, surfaces, control law, and compositions for all of the variants, is thoroughly provided. This study concludes that the SMC and its improved variants can perform satisfactorily under parametric and uncertain scenarios. The simple SMC suffers from an inherent chattering problem, which can be mitigated by the high-order and fractional-order control schemes discussed in this study. A tabulated investigation presents the techniques proposed in the literature with the error, surfaces, WECS type, inverter type, and their findings. This will help the researcher in finding appropriate surfaces and control types according to their problem. Finally, a numerical and mathematical comparision of IOSMC and FOSMC techniques was carried out to validate the effectiveness of FOSMC techniques in WECS. The results and discussion declare that the inherent problem of chattering in SMC can be overcome by using fractional-order techniques. Moreover, the results also declare that the composite SMC improved the performance of the IOSMC and FOSMC scheme. The future directions and challenges in the field of SMC for WECS can be summarized as follows:
The main problem in sliding mode control is the chattering phenomenon. The WECS is a highly complex system that can be adversely affected by the compromise between the steady state error and level of chattering. HOSMC, adaptive systems, and continuous SMC can be further enhanced to cope with this problem. The discretization of SMC schemes for digital implementation is also a challenging problem, as the active features of SMC, which includes robustness, ideal motion, and stability, are valid in the continuous domain. The fractional-order control schemes show a good performance but the high computational complexity as validated in Figure 18 is a highlighted problem. The exact approximation of fractional-order operators needs investigation in order to further minimize the IOSMC problems. The introduction of fuzzy control theory and neural networks to the SMC schemes can improve the performance of SMC in terms of chattering reduction and robustness. The lack of robustness in high-order SMC can also be improved using artificial neural networks [221] and fuzzy logic control theory.

Author Contributions

Conceptualization, I.S.; methodology, I.S.; software, I.S.; validation, S.U.; resources, funding, supervision, J.-S.R.; review, A.A.-D.; writing, review, L.K.; Writing—original draft preparation, I.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (No. NRF-2022R1A2C2004874). This work was supported by the Human Resources Development (No.20204030200090) of the Korea Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Korea government Ministry of Trade, Industry and Energy. This work was supported 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. 20214000000280).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. IEA. World Energy Balances; OECD Publishing: Paris, France, 2018; Available online: https://www.oecd-ilibrary.org/energy/world-energy-balances-2018_world_energy_bal-2018-en (accessed on 15 January 2022).
  2. Sami, I.; Ullah, S.; Ullah, N.; Ro, J.-S.J. Sensorless fractional order composite sliding mode control design for wind generation system. Isa Trans. 2021, 111, 275–289. [Google Scholar] [CrossRef] [PubMed]
  3. Nazir, M.S.; Ali, N.; Bilal, M.; Iqbal, H.M.J. Potential environmental impacts of wind energy development: A global perspective. Curr. Opin. Environ. Sci. Health 2020, 13, 85–90. [Google Scholar] [CrossRef]
  4. Sonderegger, R.C.; Henderson, D.; Bubb, S.; Steury, J.J. Distributed asset insight. IEEE Power Energy Mag. 2004, 2, 32–39. [Google Scholar] [CrossRef]
  5. Puttgen, H.B.; Macgregor, P.R.; Lambert, F.C.J. Distributed generation: Semantic hype or the dawn of a new era? IEEE Power Energy Mag. 2003, 1, 22–29. [Google Scholar] [CrossRef]
  6. Rahman, S.J. Green power: What is it and where can we find it? IEEE Power Energy Mag. 2003, 1, 30–37. [Google Scholar] [CrossRef]
  7. Martin, G.J. Renewable energy gets the “green” light in Chicago. IEEE Power Energy Mag. 2003, 1, 34–39. [Google Scholar] [CrossRef]
  8. Kumar, D.; Chatterjee, K.J. A review of conventional and advanced MPPT algorithms for wind energy systems. Renew. Sustain. Energy Rev. 2016, 55, 957–970. [Google Scholar] [CrossRef]
  9. Fernando, D.B.; Hernán, d.B.; Ricardo, J.M. Wind Turbine Control Systems: Principles, Modelling and Gain Scheduling Design; advances in industrial control; Springer: London, UK, 2006. [Google Scholar]
  10. Xu, F.; Zhang, J.; Cheng, M. Analysis of double objectives control for wind power generation system with frequency separation. In Proceedings of the 2011 4th International Conference on Electric Utility Deregulation and Restructuring and Power Technologies (DRPT), Weihai, China, 6–9 July 2011; pp. 1366–1371. [Google Scholar]
  11. Abad, G.; Lopez, J.; Rodriguez, M.; Marroyo, L.; Iwanski, G. Doubly Fed Induction Machine: Modeling and Control for Wind Energy Generation; John Wiley & Sons: New York, NY, USA, 2011. [Google Scholar]
  12. Baroudi, J.A.; Dinavahi, V.; Knight, A.M. A review of power converter topologies for wind generators. Renew. Energy 2007, 32, 2369–2385. [Google Scholar] [CrossRef]
  13. Iglesias, R.L.; Arantegui, R.L.; Alonso, M.A.J. Power electronics evolution in wind turbines—A market-based analysis. Renew. Sustain. Energy Rev. 2011, 15, 4982–4993. [Google Scholar] [CrossRef]
  14. Chakraborty, A. Advancements in power electronics and drives in interface with growing renewable energy resources. Renew. Sustain. Energy Rev. 2011, 15, 1816–1827. [Google Scholar] [CrossRef]
  15. Boldea, I. Wound Rotor Induction Generators (Wrigs). Variable Speed Generators. Politehnica University of Timisoara; CRC Press: Boca Raton, FL, USA, 2015; pp. 1–36. [Google Scholar]
  16. Li, H.; Chen, Z.J. Overview of different wind generator systems and their comparisons. IET Renew. Power Gener. 2008, 2, 123–138. [Google Scholar] [CrossRef]
  17. Bellarmine, G.T.; Urquhart, J. Wind energy for the 1990s and beyond. Energy Convers. Manag. 1996, 37, 1741–1752. [Google Scholar] [CrossRef]
  18. Sule, A.H.; Mokhtar, A.S.B.; Jamian, J.J.B. Optimal PI Pitch Control of SCIG Wind Turbine Using Grey Wolf Optimizer for Dynamic Stability. Nov. Perspect. Eng. Res. 2022, 7, 45–68. [Google Scholar]
  19. Zribi, M.; Alrifai, M.; Rayan, M.J. Sliding mode control of a variable-speed wind energy conversion system using a squirrel cage induction generator. Energies 2017, 10, 604. [Google Scholar] [CrossRef]
  20. Naik, K.; Gupta, C. Fuzzy logic based pitch angle controller/or SCIG based wind energy system. In Proceedings of the 2017 Recent Developments in Control, Automation & Power Engineering (RDCAPE), Noida, India, 26–27 October 2017; pp. 60–65. [Google Scholar]
  21. Parida, A.; Chatterjee, D.J. Integrated DFIG–SCIG-based wind energy conversion system equipped with improved power generation capability. IET Gener. Transm. Distrib. 2017, 11, 3791–3800. [Google Scholar] [CrossRef]
  22. Torrey, D.A. Variable-reluctance generators in wind-energy systems. In Proceedings of the IEEE Power Electronics Specialist Conference-PESC’93, Seattle, WA, USA, 20–24 June 1993; pp. 561–567. [Google Scholar]
  23. Arifin, A.; Al-Bahadly, I. Switched reluctance generator for variable speed wind energy applications. Smart Grid Renew. Energy 2011, 2, 27. [Google Scholar] [CrossRef]
  24. Xia, Y.; Ahmed, K.H.; Williams, B.W.J. A new maximum power point tracking technique for permanent magnet synchronous generator based wind energy conversion system. IEEE Trans. Power Electron. 2011, 26, 3609–3620. [Google Scholar] [CrossRef]
  25. Rolan, A.; Luna, A.; Vazquez, G.; Aguilar, D.; Azevedo, G. Modeling of a variable speed wind turbine with a permanent magnet synchronous generator. In Proceedings of the 2009 IEEE International Symposium on Industrial Electronics, Seoul, Korea, 5–8 July 2009; pp. 734–739. [Google Scholar]
  26. Levy, D.J. Stand alone induction generators. Electr. Power Syst. Res. 1997, 41, 191–201. [Google Scholar] [CrossRef]
  27. Blaabjerg, F.; Ma, K. Future on power electronics for wind turbine systems. IEEE J. Emerg. Sel. Top. Power Electron. 2013, 1, 139–152. [Google Scholar] [CrossRef]
  28. Hatziadoniu, C.; Nikolov, E.; Pourboghrat, F.J. Power conditioner control and protection for distributed generators and storage. IEEE Trans. Power Syst. 2003, 18, 83–90. [Google Scholar] [CrossRef]
  29. Santelo, T.N.; de Oliveira, C.M.R.; Maciel, C.D.; de A Monteiro, J.R.B. Wind Turbine Failures Review and Trends. J. Control. Autom. Electr. Syst. 2022, 33, 505–521. [Google Scholar] [CrossRef]
  30. Hansen, A.D.; Iov, F.; Blaabjerg, F.; Hansen, L.H.J. Review of contemporary wind turbine concepts and their market penetration. Wind. Eng. 2004, 28, 247–263. [Google Scholar] [CrossRef]
  31. Chinchilla, M.; Arnaltes, S.; Burgos, J.C. Control of permanent-magnet generators applied to variable-speed wind-energy systems connected to the grid. IEEE Trans. Energy Convers. 2006, 21, 130–135. [Google Scholar] [CrossRef]
  32. Yamamura, N.; Ishida, M.; Hori, T. A simple wind power generating system with permanent magnet type synchronous generator. In Proceedings of the IEEE 1999 International Conference on Power Electronics and Drive Systems. PEDS’99 (Cat. No. 99TH8475), Hong Kong, China, 27–29 July 1999; Volume 2, pp. 849–854. [Google Scholar]
  33. Muller, S.; Deicke, M.; Doncker, R.W.J.D. Doubly fed induction generator systems for wind turbines. IEEE Ind. Appl. Mag. 2002, 8, 26–33. [Google Scholar] [CrossRef]
  34. Santos-Martin, D.; Rodriguez-Amenedo, J.L.; Arnaltes, S.J. Providing ride-through capability to a doubly fed induction generator under unbalanced voltage dips. IEEE Trans. Power Electron. 2009, 24, 1747–1757. [Google Scholar] [CrossRef]
  35. Blaabjerg, F.; Chen, Z.; Kjaer, S.B. Power electronics as efficient interface in dispersed power generation systems. IEEE Trans. Power Electron. 2004, 19, 1184–1194. [Google Scholar] [CrossRef]
  36. Sami, I.; Ullah, N.; Muyeen, S.; Techato, K.; Chowdhury, M.S.; Ro, J.-S.J. Control methods for standalone and grid connected micro-hydro power plants with synthetic inertia frequency support: A comprehensive review. IEEE Access 2020, 8, 176313–176329. [Google Scholar] [CrossRef]
  37. Islam, M.R.; Guo, Y.; Zhu, J.J. A review of offshore wind turbine nacelle: Technical challenges, and research and developmental trends. Renew. Sustain. Energy Rev. 2014, 33, 161–176. [Google Scholar] [CrossRef]
  38. Zhu, Z.; Hu, J.J. Electrical machines and power-electronic systems for high-power wind energy generation applications: Part II–power electronics and control systems. Compel- Int. J. Comput. Math. Electr. Electron. Eng. 2013, 32. [Google Scholar] [CrossRef]
  39. Melício, R.; Mendes, V.; Catalão, J.P.d.S.J. Fractional-order control and simulation of wind energy systems with PMSG/full-power converter topology. Energy Convers. Manag. 2010, 51, 1250–1258. [Google Scholar] [CrossRef]
  40. Slootweg, J.; Haan, S.D.; Polinder, H.; Kling, W.J. General model for representing variable speed wind turbines in power system dynamics simulations. IEEE Trans. Power Syst. 2010, 18, 144–151. [Google Scholar] [CrossRef]
  41. Chen, Z.; Spooner, E. Grid power quality with variable speed wind turbines. IEEE Trans. Energy Convers. 2001, 16, 148–154. [Google Scholar] [CrossRef]
  42. Yazdani, A.; Iravani, R.J. A neutral-point clamped converter system for direct-drive variable-speed wind power unit. IEEE Trans. Energy Convers. 2006, 21, 596–607. [Google Scholar] [CrossRef]
  43. Carrasco, J.M.; Franquelo, L.G.; Bialasiewicz, J.T.; Galván, E.; PortilloGuisado, R.C.; Prats, M.M.; Moreno-Alfonso, N. Power-electronic systems for the grid integration of renewable energy sources: A survey. IEEE Trans. Ind. Electron. 2006, 53, 1002–1016. [Google Scholar] [CrossRef]
  44. Chen, Z.; Guerrero, J.M.; Blaabjerg, F. A review of the state of the art of power electronics for wind turbines. IEEE Trans. Power Electron. 2009, 24, 1859–1875. [Google Scholar] [CrossRef]
  45. Amei, K.; Takayasu, Y.; Ohji, T.; Sakui, M. A maximum power control of wind generator system using a permanent magnet synchronous generator and a boost chopper circuit. In Proceedings of the Power Conversion Conference-Osaka 2002 (Cat. No. 02TH8579), Osaka, Japan, 2–5 April 2002; Volume 3, pp. 1447–1452. [Google Scholar]
  46. Datta, R.; Ranganathan, V. Variable-speed wind power generation using doubly fed wound rotor induction machine-a comparison with alternative schemes. IEEE Trans. Energy Convers. 2002, 17, 414–421. [Google Scholar] [CrossRef]
  47. Suzuki, T.; Okitsu, H.; Kawahito, T. Characteristics of a small wind-power system with DC generator. IEE Proc. -Electr. Power Appl. 1982, 129, 217–220. [Google Scholar] [CrossRef]
  48. Yang, G.; Li, H. Design and analysis of a newly brushless DC wind generator. In Proceedings of the 2008 World Automation Congress, Waikoloa, HI, USA, 28 September–2 October 2008; pp. 1–5. [Google Scholar]
  49. Buehring, I.; Freris, L. Control policies for wind-energy conversion systems. IET J. 1981, 128, 253–261. [Google Scholar] [CrossRef]
  50. Abdullah, M.A.; Yatim, A.; Tan, C.W.; Saidur, R.J. A review of maximum power point tracking algorithms for wind energy systems. Renew. Sustain. Energy Rev. 2012, 16, 3220–3227. [Google Scholar] [CrossRef]
  51. Kooning, J.D.D.; Gevaert, L.; de Vyver, J.V.; Vandoorn, T.L.; Vandevelde, L. Online estimation of the power coefficient versus tip-speed ratio curve of wind turbines. In Proceedings of the IECON 2013-39th Annual Conference of the IEEE Industrial Electronics Society, Vienna, Austria, 10–13 November 2013; pp. 1792–1797. [Google Scholar]
  52. Knight, A.M.; Peters, G.E.J. Simple wind energy controller for an expanded operating range. IEEE Trans. Energy Convers. 2005, 20, 459–466. [Google Scholar] [CrossRef]
  53. Nasiri, M.; Milimonfared, J.; Fathi, S.J. Modeling, analysis and comparison of TSR and OTC methods for MPPT and power smoothing in permanent magnet synchronous generator-based wind turbines. Energy Convers. Manag. 2014, 86, 892–900. [Google Scholar] [CrossRef]
  54. Yaramasu, V.; Wu, B. Basics of Wind Energy Conversion Systems (Wecs); Wiley-IEEE Press: New York, NY, USA, 2017. [Google Scholar]
  55. Tan, K.; Islam, S.J. Optimum control strategies in energy conversion of PMSG wind turbine system without mechanical sensors. IEEE Trans. Energy Convers. 2004, 19, 392–399. [Google Scholar] [CrossRef]
  56. Ganjefar, S.; Ghassemi, A.A.; Ahmadi, M.M.J. Improving efficiency of two-type maximum power point tracking methods of tip-speed ratio and optimum torque in wind turbine system using a quantum neural network. Energy Convers. Manag. 2014, 67, 444–453. [Google Scholar] [CrossRef]
  57. Agarwal, V.; Aggarwal, R.K.; Patidar, P.; Patki, C.J. A novel scheme for rapid tracking of maximum power point in wind energy generation systems. IEEE Trans. Energy Convers. 2009, 25, 228–236. [Google Scholar] [CrossRef]
  58. Yan, R.; Saha, T.K.J. A new tool to estimate maximum wind power penetration level: In perspective of frequency response adequacy. Appl. Energy 2015, 154, 209–220. [Google Scholar]
  59. Wei, C.; Zhang, Z.; Qiao, W.; Qu, L.J. Reinforcement-learning-based intelligent maximum power point tracking control for wind energy conversion systems. IEEE Trans. Ind. Electron. 2015, 62, 6360–6370. [Google Scholar] [CrossRef]
  60. Simoes, M.G.; Bose, B.K.; Spiegel, R.J.J. Fuzzy logic based intelligent control of a variable speed cage machine wind generation system. IEEE Trans. Power Electron. 1997, 12, 87–95. [Google Scholar] [CrossRef]
  61. Hilloowala, R.M.; Sharaf, A.M. A rule-based fuzzy logic controller for a PWM inverter in a stand alone wind energy conversion scheme. In Proceedings of the Conference Record of the 1993 IEEE Industry Applications Conference Twenty-Eighth IAS Annual Meeting, Toronto, ON, Canada, 2–8 October 1993; pp. 2066–2073. [Google Scholar]
  62. Chen, Z.; Gomez, S.A.; McCormick, M. A fuzzy logic controlled power electronic system for variable speed wind energy conversion systems. In Proceedings of the 8th International Conference on Power Electronics and Variable Speed Drives, London, UK, 18–19 September 2000. [Google Scholar]
  63. Mohamed, A.Z.; Eskander, M.N.; Ghali, F.A.J. Fuzzy logic control based maximum power tracking of a wind energy system. Renew. Energy 2001, 23, 235–245. [Google Scholar] [CrossRef]
  64. Kumar, G.B. Optimal power point tracking of solar and wind energy in a hybrid wind solar energy system. Int. J. Energy Environ. Eng. 2022, 13, 77–103. [Google Scholar] [CrossRef]
  65. Morimoto, S.; Nakayama, H.; Sanada, M.; Takeda, Y.J. Sensorless output maximization control for variable-speed wind generation system using IPMSG. IEEE Trans. Ind. Appl. 2005, 41, 60–67. [Google Scholar] [CrossRef]
  66. Hui, J.; Bakhshai, A.; Jain, P.K. An adaptive approximation method for maximum power point tracking (MPPT) in wind energy systems. In Proceedings of the 2011 IEEE Energy Conversion Congress and Exposition, Phoenix, AZ, USA, 17–22 September 2011; pp. 2664–2669. [Google Scholar]
  67. Yessef, M.; Bossoufi, B.; Taoussi, M.; Lagrioui, A.; Chojaa, H. Overview of control strategies for wind turbines: ANNC, FLC, SMC, BSC, and PI controllers. Wind Energy 2022. [Google Scholar] [CrossRef]
  68. Noussi, K.; Abouloifa, A.; Katir, H.; Lachkar, I.; Aouadi, C.; Aourir, M.; Otmani, F.E. Integral Backstepping Control Based on High Gain Observer for DFIG-Based Wind Energy Conversion System. IFAC-PapersOnLine 2022, 55, 653–658. [Google Scholar] [CrossRef]
  69. Civelek, Z.; Lüy, M.; Çam, E.; Mamur, H. A new fuzzy logic proportional controller approach applied to individual pitch angle for wind turbine load mitigation. Renew. Energy 2017, 111, 708–717. [Google Scholar] [CrossRef]
  70. Medjber, A.; Guessoum, A.; Belmili, H.; Mellit, A.J. New neural network and fuzzy logic controllers to monitor maximum power for wind energy conversion system. Energy 2016, 106, 137–146. [Google Scholar] [CrossRef]
  71. Lin, W.-M.; Hong, C.-M.; Cheng, F.-S.J. Fuzzy neural network output maximization control for sensorless wind energy conversion system. Energy 2010, 35, 592–601. [Google Scholar] [CrossRef]
  72. Pichan, M.; Rastegar, H.; Monfared, M.J. Two fuzzy-based direct power control strategies for doubly-fed induction generators in wind energy conversion systems. Energy 2013, 51, 154–162. [Google Scholar] [CrossRef]
  73. Yang, B.; Jiang, L.; Wang, L.; Yao, W.; Wu, Q.J. Nonlinear maximum power point tracking control and modal analysis of DFIG based wind turbine. Int. J. Electr. Power Energy Syst. 2016, 74, 429–436. [Google Scholar] [CrossRef]
  74. Liu, X.; Kong, X.J. Nonlinear model predictive control for DFIG-based wind power generation. IEEE Trans. Autom. Sci. Eng. 2013, 11, 1046–1055. [Google Scholar] [CrossRef]
  75. Gani, A.; Mohammadi, K.; Shamshirband, S.; Altameem, T.A.; Petković, D.J. A combined method to estimate wind speed distribution based on integrating the support vector machine with firefly algorithm. Environ. Prog. Sustain. Energy 2016, 35, 867–875. [Google Scholar] [CrossRef]
  76. Chen, P.; Han, D.; Li, K.-C. Robust Adaptive Control of Maximum Power Point Tracking for Wind Power System. IEEE Access 2020, 8, 214538–214550. [Google Scholar] [CrossRef]
  77. Cao, Z.; Niu, Y.; Karimi, H.R.J.I.S. Sliding mode control of automotive electronic valve system under weighted try-once-discard protocol. Inf. Sci. 2020, 515, 324–340. [Google Scholar] [CrossRef]
  78. Surinkaew, T.; Ngamroo, I.J. Robust power oscillation damper design for DFIG-based wind turbine based on specified structure mixed H2/H∞ control. Renew. Energy 2014, 66, 15–24. [Google Scholar] [CrossRef]
  79. Das, S.; Subudhi, B.J. A robust active and reactive power control scheme with multiloop disturbance rejection for a wind energy conversion system. IEEE Trans. Sustain. Energy 2018, 10, 1664–1671. [Google Scholar] [CrossRef]
  80. Ghanbarpour, K.; Bayat, F.; Jalilvand, A.J. Dependable power extraction in wind turbines using model predictive fault tolerant control. Int. J. Electr. PowerEnergy Syst. 2020, 118, 105802. [Google Scholar] [CrossRef]
  81. Kühne, P.; Pöschke, F.; Schulte, H.J. Fault estimation and fault-tolerant control of the FAST NREL 5-MW reference wind turbine using a proportional multi-integral observer. Int. J. Adapt. Control. Process. 2018, 32, 568–585. [Google Scholar] [CrossRef]
  82. Aguilar, M.E.B.; Coury, D.V.; Reginatto, R.; Monaro, R.M.J. Multi-objective PSO applied to PI control of DFIG wind turbine under electrical fault conditions. Electr. Power Syst. Res. 2020, 180, 106081. [Google Scholar] [CrossRef]
  83. Shaker, M.S.; Patton, R.J.J. Active sensor fault tolerant output feedback tracking control for wind turbine systems via T–S model. Eng. Appl. Artif. Intell. 2014, 34, 1–12. [Google Scholar] [CrossRef]
  84. Kamal, E.; Aitouche, A.; Oueidat, M.J. Fuzzy fault-tolerant control of wind-diesel hybrid systems subject to sensor faults. IEEE Trans. Sustain. Energy 2013, 4, 857–866. [Google Scholar] [CrossRef]
  85. Alqahtani, A.H.; Utkin, V.I. Self-optimization of photovoltaic system power generation based on sliding mode control. In Proceedings of the IECON 2012-38th Annual Conference on IEEE Industrial Electronics Society, Montreal, QC, Canada, 25–28 October 2012; pp. 3468–3474. [Google Scholar]
  86. Levron, Y.; Shmilovitz, D.J. Maximum power point tracking employing sliding mode control. IEEE Trans. Circuits Syst. Regul. Pap. 2013, 60, 724–732. [Google Scholar] [CrossRef]
  87. Costa, J.P.D.; Pinheiro, H.; Degner, T.; Arnold, G. Robust controller for DFIGs of grid-connected wind turbines. IEEE Trans. Ind. Electron. 2010, 58, 4023–4038. [Google Scholar] [CrossRef]
  88. Pande, V.; Mate, U.; Kurode, S.J. Discrete sliding mode control strategy for direct real and reactive power regulation of wind driven DFIG. Electr. Power Syst. Res. 2013, 100, 73–81. [Google Scholar] [CrossRef]
  89. Majout, B.; Bossoufi, B.; Bouderbala, M.; Masud, M.; Al-Amri, J.F.; Taoussi, M.; El Mahfoud, M.; Motahhir, S.; Karim, M. Improvement of PMSG-Based Wind Energy Conversion System Using Developed Sliding Mode Control. Energies 2022, 15, 1625. [Google Scholar] [CrossRef]
  90. Ardjoun, S.A.E.M.; Abid, M. Fuzzy sliding mode control applied to a doubly fed induction generator for wind turbines. Turk. J. Electr. Eng. Comput. Sci. 2015, 23, 1673–1686. [Google Scholar] [CrossRef]
  91. Djoudi, A.; Chekireb, H.; Bacha, S.J. Low-cost sliding mode control of WECS based on DFIG with stability analysis. Turk. J. Electr. Eng. Comput. Sci. 2015, 23, 1698–1714. [Google Scholar] [CrossRef]
  92. Utkin, V.; Lee, H. Chattering problem in sliding mode control systems. In Proceedings of the VSS’06: International Workshop on Variable Structure Systems, Alghero, Italy, 5–7 June 2006; pp. 346–350. [Google Scholar]
  93. Kassem, A.M.; Hasaneen, K.M.; Yousef, A.M.J. Dynamic modeling and robust power control of DFIG driven by wind turbine at infinite grid. Int. J. Electr. Power Energy Syst. 2013, 44, 375–382. [Google Scholar] [CrossRef]
  94. Laghrouche, S.; Plestan, F.; Glumineau, A.J. Higher order sliding mode control based on integral sliding mode. Automatica 2007, 43, 531–537. [Google Scholar] [CrossRef]
  95. Zholtayev, D.; Rubagotti, M.; Do, T.D. Adaptive super-twisting sliding mode control for maximum power point tracking of PMSG-based wind energy conversion systems. Renew. Energy 2022, 183, 877–889. [Google Scholar] [CrossRef]
  96. Mazouz, F.; Belkacem, S.; Boukhalfa, G. Second Order Sliding Mode Control Based DPC of DFIG Using SVM. In Proceedings of the International Conference on Artificial Intelligence in Renewable Energetic Systems, Tipasa, Algeriapp, 24–26 November 2021; Springer: Cham, Siwtzerland, 2021; pp. 743–750. [Google Scholar]
  97. Dekali, Z.; Baghli, L.; Boumediene, A. Improved Super Twisting Based High Order Direct Power Sliding Mode Control of a Connected DFIG Variable Speed Wind Turbine. Period. Polytech. Electr. Eng. Comput. Sci. 2021. [Google Scholar] [CrossRef]
  98. Abdeddaim, S.; Betka, A. Optimal tracking and robust power control of the DFIG wind turbine. Int. J. Electr. Power Energy Syst. 2013, 49, 234–242. [Google Scholar] [CrossRef]
  99. Du, C.; Han, Y.; Li, S. A Barrier Function-Based Variable-Gain SOSM Power Control Scheme for DFIG Wind Turbine. In Proceedings of the 2021 Chinese Intelligent Automation Conference, Zhanjiang, China, 5–7 November 2021; Springer: Singapore, 2022; pp. 116–123. [Google Scholar]
  100. Hwang, S.; Park, J.B.; Joo, Y.H.J. Disturbance observer-based integral fuzzy sliding-mode control and its application to wind turbine system. IET Control. Theory Appl. 2019, 13, 1891–1900. [Google Scholar] [CrossRef]
  101. Song, Z.; Shi, T.; Xia, C.; Chen, W.J. A novel adaptive control scheme for dynamic performance improvement of DFIG-Based wind turbines. Energy 2012, 38, 104–117. [Google Scholar] [CrossRef]
  102. Ebrahimkhani, S.J. Robust fractional order sliding mode control of doubly-fed induction generator (DFIG)-based wind turbines. ISA Trans. 2016, 63, 343–354. [Google Scholar] [CrossRef] [PubMed]
  103. Xiong, L.; Li, P.; Ma, M.; Wang, Z.; Wang, J.J. Output power quality enhancement of PMSG with fractional order sliding mode control. Int. J. Electr. Power Energy Syst. 2020, 115, 105402. [Google Scholar] [CrossRef]
  104. Ullah, N.; Ali, M.A.; Ibeas, A.; Herrera, J.J. Adaptive fractional order terminal sliding mode control of a doubly fed induction generator-based wind energy system. IEEE Access 2017, 5, 21368–21381. [Google Scholar] [CrossRef]
  105. Sami, I.; Ullah, S.; Ali, Z.; Ullah, N.; Ro, J.-S.J. A Super Twisting Fractional Order Terminal Sliding Mode Control for DFIG-Based Wind Energy Conversion System. Energies 2020, 13, 2158. [Google Scholar] [CrossRef]
  106. Ullah, N.; Sami, I.; Chowdhury, M.S.; Techato, K.; Alkhammash, H.J. Artificial Intelligence Integrated Fractional Order Control of Doubly Fed Induction Generator Based Wind Energy System. IEEE Access 2020, 9, 5734–5748. [Google Scholar] [CrossRef]
  107. Tamaarat, A.; Benakcha, A.J. Performance of PI controller for control of active and reactive power in DFIG operating in a grid-connected variable speed wind energy conversion system. Front. Energy 2014, 8, 371–378. [Google Scholar] [CrossRef]
  108. Bossoufi, B.; Karim, M.; Lagrioui, A.; Taoussi, M.; Derouich, A.J. Observer backstepping control of DFIG-Generators for wind turbines variable-speed: FPGA-based implementation. Renew. Energy 2015, 81, 903–917. [Google Scholar] [CrossRef]
  109. Taraft, S.; Rekioua, D.; Aouzellag, D.; Bacha, S.J. A proposed strategy for power optimization of a wind energy conversion system connected to the grid. Energy Convers. Manag. 2015, 101, 489–502. [Google Scholar] [CrossRef]
  110. Hu, J.; Nian, H.; Hu, B.; He, Y.; Zhu, Z.J. Direct active and reactive power regulation of DFIG using sliding-mode control approach. IEEE Trans. Energy Convers. 2010, 25, 1028–1039. [Google Scholar] [CrossRef]
  111. Fallaha, C.J.; Saad, M.; Kanaan, H.Y.; Al-Haddad, K.J. Sliding-mode robot control with exponential reaching law. IEEE Trans. Ind. Electron. 2010, 58, 600–610. [Google Scholar] [CrossRef]
  112. Wang, J.; Bo, D.; Ma, X.; Zhang, Y.; Li, Z.; Miao, Q.J. Adaptive back-stepping control for a permanent magnet synchronous generator wind energy conversion system. Int. J. Hydrogen Energy 2019, 44, 3240–3249. [Google Scholar] [CrossRef]
  113. Shihabudheen, K.; Raju, S.K.; Pillai, G.N.J. Control for grid-connected DFIG-based wind energy system using adaptive neuro-fuzzy technique. Int. Trans. Electr. Energy Syst. 2018, 28, e2526. [Google Scholar] [CrossRef]
  114. Mérida, J.; Aguilar, L.T.; Dávila, J.J. Analysis and synthesis of sliding mode control for large scale variable speed wind turbine for power optimization. Renew. Energy 2014, 71, 715–728. [Google Scholar] [CrossRef]
  115. Xiong, L.; Li, P.; Wu, F.; Ma, M.; Khan, M.W.; Wang, J.J. A coordinated high-order sliding mode control of DFIG wind turbine for power optimization and grid synchronization. Int. J. Electr. Power Energy Syst. 2019, 105, 679–689. [Google Scholar] [CrossRef]
  116. Beltran, B.; Benbouzid, M.E.H.; Ahmed-Ali, T.J. Second-order sliding mode control of a doubly fed induction generator driven wind turbine. IEEE Trans. Energy Convers. 2012, 27, 261–269. [Google Scholar] [CrossRef]
  117. Beltran, B.; Ahmed-Ali, T.; Benbouzid, M.E.H.J. High-order sliding-mode control of variable-speed wind turbines. IEEE Trans. Ind. Electron. 2008, 56, 3314–3321. [Google Scholar] [CrossRef]
  118. Zhang, C.; Gutierrez, S.; Plestan, F.; de León-Morales, J.J. Adaptive super-twisting control of floating wind turbines with collective blade pitch control. IFAC-PapersOnLine 2019, 52, 117–122. [Google Scholar] [CrossRef]
  119. Chen, H.; Li, Q.; Tang, S.; Nadia, A.A.; Han, J.; Wang, T.; Benbouzid, M. Adaptive super-twisting control of doubly salient permanent magnet generator for tidal stream turbine. Int. J. Electr. Power Energy Syst. 2021, 128, 106772. [Google Scholar] [CrossRef]
  120. Milosavljevic, D. General conditions for existence of a quasi-sliding mode on the switching hyperplane in discrete variable structure systems. Autom. Remote Control 1985, 46, 307–314. [Google Scholar]
  121. Bartoszewicz, A. Discrete-time quasi-sliding-mode control strategies. IEEE Trans. Ind. Electron. 1998, 45, 633–637. [Google Scholar] [CrossRef]
  122. Ngadungon, R.B.N.; Sam, Y.B.M. Discrete time sliding mode control using multirate output feedback to reduce chattering. In Proceedings of the 2010 IEEE Student Conference on Research and Development (SCOReD), Kuala Lumpur, Malaysia, 13–14 December 2010; pp. 236–239. [Google Scholar]
  123. Beltran, B.; Ahmed-Ali, T.; Benbouzid, M.E.H. Sliding mode power control of variable-speed wind energy conversion systems. IEEE Trans. Energy Convers. 2008, 23, 551–558. [Google Scholar] [CrossRef]
  124. Mozayan, S.M.; Saad, M.; Vahedi, H.; Fortin-Blanchette, H.; Soltani, M.J. Sliding mode control of PMSG wind turbine based on enhanced exponential reaching law. IEEE Trans. Ind. Electron. 2016, 63, 6148–6159. [Google Scholar] [CrossRef]
  125. Hung, J.Y.; Gao, W.; Hung, J.C. Variable structure control: A survey. IEEE Trans. Ind. Electron. 1993, 40, 2–22. [Google Scholar] [CrossRef]
  126. Kelkoul, B.; Boumediene, A.J. Stability analysis and study between classical sliding mode control (SMC) and super twisting algorithm (STA) for doubly fed induction generator (DFIG) under wind turbine. Energy Convers. Manag. 2020, 214, 118871. [Google Scholar] [CrossRef]
  127. Fridman, L.; Moreno, J.A.; Bandyopadhyay, B.; Kamal, S.; Chalanga, A. Continuous nested algorithms: The fifth generation of sliding mode controllers. In Recent Advances in Sliding Modes: From Control to Intelligent Mechatronics; Springer: Cham, Switzerland, 2015; pp. 5–35. [Google Scholar]
  128. Bekakra, Y.; Attous, D.B. Sliding mode controls of active and reactive power of a DFIG with MPPT for variable speed wind energy conversion. Aust. J. Basic Appl. Sci. 2011, 5, 2274–2286. [Google Scholar]
  129. Machmoum, M.; Poitiers, F. Sliding mode control of a variable speed wind energy conversion system with DFIG. In Proceedings of the International Conference and Exhibition on Ecologic Vehicles and Renewable Energies, Monte-Carlo, Monaco, 8–10 May 2019; pp. 26–29. [Google Scholar]
  130. Liu, X.; Han, Y.; Wang, C.J. Second-order sliding mode control for power optimisation of DFIG-based variable speed wind turbine. IET Renew. Power Gener. 2016, 11, 408–418. [Google Scholar] [CrossRef]
  131. Beltran, B.; Benbouzid, M.; Ahmed-Ali, T. High-order sliding mode control of a DFIG-based wind turbine for power maximization and grid fault tolerance. In Proceedings of the 2009 IEEE International Electric Machines and Drives Conference, Miami, FL, USA, 3–6 May 2009; pp. 183–189. [Google Scholar]
  132. Levant, A.J. Higher-order sliding modes, differentiation and output-feedback control. Int. J. Control 2003, 76, 924–941. [Google Scholar] [CrossRef]
  133. Benelghali, S.; Benbouzid, M.; Charpentier, J.; Ahmed-Ali, T.; Munteanu, I. Experimental validation of a marine current turbine simulator: Application to a PMSG-based system second-order sliding mode control. IEEE Trans. Ind. Electron. 2011, 58, 118–126. [Google Scholar] [CrossRef]
  134. Elghali, S.E.B.; Benbouzid, M.E.H.; Ahmed-Ali, T.; Charpentier, J.F.J. High-order sliding mode control of a marine current turbine driven doubly-fed induction generator. IEEE J. Ocean. Eng. 2010, 35, 402–411. [Google Scholar] [CrossRef]
  135. Benbouzid, M.; Beltran, B.; Amirat, Y.; Yao, G.; Han, J.; Mangel, H. Second-order sliding mode control for DFIG-based wind turbines fault ride-through capability enhancement. ISA Trans. 2014, 53, 827–833. [Google Scholar] [CrossRef] [PubMed]
  136. Dursun, E.H.; Kulaksiz, A.A.J. Second-order sliding mode voltage-regulator for improving MPPT efficiency of PMSG-based WECS. Int. J. Electr. Power Energy Syst. 2020, 121, 106149. [Google Scholar] [CrossRef]
  137. Wang, J.; Bo, D.; Miao, Q.; Li, Z.; Wu, X.; Lv, D.J. Maximum power point tracking control for a doubly fed induction generator wind energy conversion system based on multivariable adaptive super-twisting approach. Int. J. Electr. Power Energy Syst. 2021, 124, 106347. [Google Scholar] [CrossRef]
  138. Valenciaga, F.; Puleston, P.J. High-order sliding control for a wind energy conversion system based on a permanent magnet synchronous generator. IEEE Trans. Energy Convers. 2008, 23, 860–867. [Google Scholar] [CrossRef]
  139. Evangelista, C.; Puleston, P.; Valenciaga, F.; Fridman, L.M.J. Lyapunov-designed super-twisting sliding mode control for wind energy conversion optimization. IEEE Trans. Ind. Electron. 2012, 60, 538–545. [Google Scholar] [CrossRef]
  140. Evangelista, C.; Valenciaga, F.; Puleston, P.J. Active and reactive power control for wind turbine based on a MIMO 2-sliding mode algorithm with variable gains. IEEE Trans. Energy Convers. 2013, 28, 682–689. [Google Scholar] [CrossRef]
  141. Merabet, A.; Ahmed, K.T.; Ibrahim, H.; Beguenane, R.J. Implementation of sliding mode control system for generator and grid sides control of wind energy conversion system. IEEE Trans. Sustain. Energy 2016, 7, 1327–1335. [Google Scholar] [CrossRef]
  142. Morshed, M.J.; Fekih, A.J. A new fault ride-through control for DFIG-based wind energy systems. Electr. Power Syst. Res. 2017, 146, 258–269. [Google Scholar] [CrossRef]
  143. Evangelista, C.A.; Pisano, A.; Puleston, P.; Usai, E.J. Receding horizon adaptive second-order sliding mode control for doubly-fed induction generator based wind turbine. IEEE Trans. Control. Syst. Technol. 2016, 25, 73–84. [Google Scholar] [CrossRef]
  144. Benamor, A.; Benchouia, M.; Srairi, K.; Benbouzid, M. A novel rooted tree optimization apply in the high order sliding mode control using super-twisting algorithm based on DTC scheme for DFIG. Int. J. Electr. Power Energy Syst. 2019, 108, 293–302. [Google Scholar] [CrossRef]
  145. Berrada, Y.; Boumhidi, I. New structure of sliding mode control for variable speed wind turbine. IFAC J. Syst. Control 2020, 14, 100113. [Google Scholar] [CrossRef]
  146. Karabacak, M.J. A new perturb and observe based higher order sliding mode MPPT control of wind turbines eliminating the rotor inertial effect. Renew. Energy 2019, 133, 807–827. [Google Scholar] [CrossRef]
  147. Xiong, L.; Li, P.; Wang, J.J. High-order sliding mode control of DFIG under unbalanced grid voltage conditions. Int. J. Electr. Power Energy Syst. 2020, 117, 105608. [Google Scholar] [CrossRef]
  148. Shtessel, Y.; Taleb, M.; Plestan, F.J. A novel adaptive-gain supertwisting sliding mode controller: Methodology and application. Automatica 2012, 48, 759–769. [Google Scholar] [CrossRef]
  149. Utkin, V.; Shi, J. Integral sliding mode in systems operating under uncertainty conditions. In Proceedings of the 35th IEEE Conference on Decision and Control, Kobe, Japan, 13 December 1996; Volume 4, pp. 4591–4596. [Google Scholar]
  150. Suleimenov, K.; Sarsembayev, B.; Phuc, B.D.H.; Do, T.D.J. Disturbance observer-based integral sliding mode control for wind energy conversion systems. Wind. Energy 2020, 23, 1026–1047. [Google Scholar] [CrossRef]
  151. Saravanakumar, R.; Jena, D.J. Validation of an integral sliding mode control for optimal control of a three blade variable speed variable pitch wind turbine. Int. J. Electr. Power Energy Syst. 2015, 69, 421–429. [Google Scholar] [CrossRef]
  152. Li, S.; Wang, H.; Tian, Y.; Aitouch, A.; Klein, J.J. Direct power control of DFIG wind turbine systems based on an intelligent proportional-integral sliding mode control. ISA Trans. 2016, 64, 431–439. [Google Scholar] [CrossRef]
  153. Chen, S.Z.; Cheung, N.C.; Wong, K.C.; Wu, J. Integral sliding-mode direct torque control of doubly-fed induction generators under unbalanced grid voltage. IEEE Trans. Energy Convers. 2009, 25, 356–368. [Google Scholar] [CrossRef]
  154. Martinez, M.I.; Susperregui, A.; Tapia, G.; Camblong, H.J. Sliding-Mode Control for a DFIG-based Wind Turbine under Unbalanced Voltage. Ifac Proc. Vol. 2011, 44, 538–543. [Google Scholar] [CrossRef]
  155. Liu, Y.; Wang, Z.; Xiong, L.; Wang, J.; Jiang, X.; Bai, G.; Liu, S. DFIG wind turbine sliding mode control with exponential reaching law under variable wind speed. Int. J. Electr. Power Energy Syst. 2018, 96, 253–260. [Google Scholar] [CrossRef]
  156. Errouissi, R.; Al-Durra, A.J. A novel PI-type sliding surface for PMSG-based wind turbine with improved transient performance. IEEE Trans. Energy Convers. 2017, 33, 834–844. [Google Scholar] [CrossRef]
  157. Sarsembayev, B.; Suleimenov, K.; Mirzagalikova, B.; Do, T.D.J. SDRE-based integral sliding mode control for wind energy conversion systems. IEEE Access 2020, 8, 51100–51113. [Google Scholar] [CrossRef]
  158. Nayeh, R.F.; Moradi, H.; Vossoughi, G.J. Multivariable robust control of a horizontal wind turbine under various operating modes and uncertainties: A comparison on sliding mode and H∞ control. Int. J. Electr. Power Energy Syst. 2020, 115, 105474. [Google Scholar] [CrossRef]
  159. Yin, X.; Jiang, Z.; Pan, L.J. Recurrent neural network based adaptive integral sliding mode power maximization control for wind power systems. Renew. Energy 2020, 145, 1149–1157. [Google Scholar] [CrossRef]
  160. Dali, A.; Abdelmalek, S.; Bakdi, A.; Bettayeb, M. A new robust control scheme: Application for MPP tracking of a PMSG-based variable-speed wind turbine. Renew. Energy 2021, 172, 1021–1034. [Google Scholar] [CrossRef]
  161. Xiong, L.; Li, J.; Li, P.; Huang, S.; Wang, Z.; Wang, J.J. Event triggered prescribed time convergence sliding mode control of DFIG with disturbance rejection capability. Int. J. Electr. Power Energy Syst. 2021, 131, 106970. [Google Scholar] [CrossRef]
  162. Yu, X.; Kaynak, O.J. Sliding-mode control with soft computing: A survey. IEEE Trans. Ind. Electron. 2009, 56, 3275–3285. [Google Scholar]
  163. Utkin, V.; Poznyak, A.; Orlov, Y.; Polyakov, A.J. Conventional and high order sliding mode control. J. Frankl. Inst. 2020, 357, 10244–10261. [Google Scholar] [CrossRef]
  164. Yu, X.; Feng, Y.; Man, Z.J. Terminal Sliding Mode Control-An Overview. IEEE Open J. Ind. Electron. Soc. 2020, 2, 36–52. [Google Scholar] [CrossRef]
  165. Zak, M.J. Terminal attractors for addressable memory in neural networks. Phys. Lett. A 1988, 133, 18–22. [Google Scholar] [CrossRef]
  166. Venkataraman, S.; Gulati, S. Control of nonlinear systems using terminal sliding modes. J. Dyn. Sys. Meas. Control. 1993, 115, 554–560. [Google Scholar] [CrossRef]
  167. Yu, X.; Man, Z.J. Model reference adaptive control systems with terminal sliding modes. Int. J. Control 1996, 64, 1165–1176. [Google Scholar] [CrossRef]
  168. Zhihong, M.; Paplinski, A.P.; Wu, H.R.J. A robust MIMO terminal sliding mode control scheme for rigid robotic manipulators. IEEE Trans. Autom. Control 1994, 39, 2464–2469. [Google Scholar] [CrossRef]
  169. Zhihong, M.; Yu, X.H.J. Terminal sliding mode control of MIMO linear systems. IEEE Trans. Circuits Syst. Fundam. Theory Appl. 1997, 44, 1065–1070. [Google Scholar] [CrossRef]
  170. Feng, Y.; Yu, X.; Man, Z.J. Non-singular terminal sliding mode control of rigid manipulators. Automatica 2002, 38, 2159–2167. [Google Scholar] [CrossRef]
  171. Yu, X.; Zhihong, M.J. Fast terminal sliding-mode control design for nonlinear dynamical systems. IEEE Trans. Circuits Syst. Fundam. Theory Appl. 2002, 49, 261–264. [Google Scholar]
  172. Yu, S.; Yu, X.; Shirinzadeh, B.; Man, Z.J. Continuous finite-time control for robotic manipulators with terminal sliding mode. Automatica 2005, 41, 1957–1964. [Google Scholar] [CrossRef]
  173. Patnaik, R.; Dash, P.J. Fast adaptive finite-time terminal sliding mode power control for the rotor side converter of the DFIG based wind energy conversion system. Sustain. Energy, Grids Netw. 2015, 1, 63–84. [Google Scholar] [CrossRef]
  174. Patnaik, R.; Dash, P.; Mahapatra, K.J. Adaptive terminal sliding mode power control of DFIG based wind energy conversion system for stability enhancement. Int. Trans. Electr. Energy Syst. 2016, 26, 750–782. [Google Scholar] [CrossRef]
  175. Yazıcı, İ.; Yaylacı, E.K.J. Discrete-time integral terminal sliding mode based maximum power point controller for the PMSG-based wind energy system. Iet Power Electron. 2019, 12, 3688–3696. [Google Scholar] [CrossRef]
  176. Zheng, X.; Feng, Y.; Han, F.; Yu, X.J. Integral-type terminal sliding-mode control for grid-side converter in wind energy conversion systems. IEEE Trans. Ind. Electron. 2018, 66, 3702–3711. [Google Scholar] [CrossRef]
  177. Morshed, M.J.; Fekih, A.J. Design of a chattering-free integral terminal sliding mode approach for DFIG-based wind energy systems. Optim. Control. Appl. Methods 2020, 41, 1718–1734. [Google Scholar] [CrossRef]
  178. Balamurugan, S.; Venkatesh, P.; Varatharajan, M. Fuzzy sliding-mode control with low pass filter to reduce chattering effect: An experimental validation on Quanser SRIP. Sādhanā 2017, 42, 1693–1703. [Google Scholar] [CrossRef]
  179. Xu, N.; Chen, Y.; Xue, A.; Zong, G.; Zhao, X.J. Event-trigger-based adaptive fuzzy hierarchical sliding mode control of uncertain under-actuated switched nonlinear systems. ISA Trans. 2019, 124, 301–310. [Google Scholar] [CrossRef]
  180. Yin, X.-x.; Lin, Y.-g.; Li, W.; Gu, Y.-j.; Lei, P.-f.; Liu, H.-w.J. Sliding mode voltage control strategy for capturing maximum wind energy based on fuzzy logic control. Int. J. Electr. Power Energy Syst. 2015, 70, 45–51. [Google Scholar] [CrossRef]
  181. Yin, X.-X.; Lin, Y.-G.; Li, W.; Liu, H.-W.; Gu, Y.-J.J. Fuzzy-logic sliding-mode control strategy for extracting maximum wind power. IEEE Trans. Energy Convers. 2015, 30, 1267–1278. [Google Scholar] [CrossRef]
  182. Do, T.D.J. Disturbance observer-based fuzzy SMC of WECSs without wind speed measurement. IEEE Access 2016, 5, 147–155. [Google Scholar] [CrossRef]
  183. Bounar, N.; Labdai, S.; Boulkroune, A. PSO–GSA based fuzzy sliding mode controller for DFIG-based wind turbine. ISA Trans. 2019, 85, 177–188. [Google Scholar] [CrossRef]
  184. Hong, C.-M.; Huang, C.-H.; Cheng, F.-S.J. Sliding mode control for variable-speed wind turbine generation systems using artificial neural network. Energy Procedia 2014, 61, 1626–1629. [Google Scholar] [CrossRef]
  185. Miller, K.S.; Ross, B. An Introduction to the Fractional Calculus and Fractional Differential Equations; Wiley: Hoboken, NJ, USA, 1993. [Google Scholar]
  186. Srivastava, H.J. Fractional Differential Equations. I Podlubny. Appl. Mech. Rev. 2000, 53, B51. [Google Scholar]
  187. Matignon, D. Stability properties for generalized fractional differential systems. In ESAIM: Proceedings; EDP Sciences: Paris, France, 1998; Volume 5, pp. 145–158. [Google Scholar]
  188. Zhang, Q.; Zhuang, X.; Liu, Y.; Wang, C.; Guo, H.J. A novel control strategy for mode seamless switching of PV converter in DC microgrid based on double integral sliding mode control. ISA Trans. 2020, 100, 469–480. [Google Scholar] [CrossRef] [PubMed]
  189. Incremona, G.P.; Rubagotti, M.; Ferrara, A.J. Sliding mode control of constrained nonlinear systems. IEEE Trans. Autom. Control. 2016, 62, 2965–2972. [Google Scholar] [CrossRef]
  190. Liu, C.; Sun, W.; Zhang, J.J. Adaptive sliding mode control for 4–wheel SBW system with Ackerman geometry. ISA Trans. 2020, 96, 103–115. [Google Scholar] [CrossRef]
  191. Chen, L.; Wang, H.; Huang, Y.; Ping, Z.; Yu, M.; Zheng, X.; Hu, Y. Robust hierarchical sliding mode control of a two-wheeled self-balancing vehicle using perturbation estimation. Mech. Syst. Process. 2020, 139, 106584. [Google Scholar] [CrossRef]
  192. Kang, S.; Yan, H.; Dong, L.; Li, C.J. Finite-time adaptive sliding mode force control for electro-hydraulic load simulator based on improved GMS friction model. Mech. Syst. Process. 2018, 102, 117–138. [Google Scholar] [CrossRef]
  193. Pradhan, S.; Singh, B.; Panigrahi, B.K.; Murshid, S.J. A composite sliding mode controller for wind power extraction in remotely located solar PV–wind hybrid system. IEEE Trans. Ind. Electron. 2018, 66, 5321–5331. [Google Scholar] [CrossRef]
  194. Yashar, M.; Bevan, G.; Kucukdemiral, I.B.; Fekih, A. Sliding mode control of wind energy conversion systems: Trends and applications. Renew. Sustain. Energy Rev. 2022, 167, 112734. [Google Scholar]
  195. Wang, H.; Shi, L.; Man, Z.; Zheng, J.; Li, S.; Yu, M.; Cao, Z. Continuous fast nonsingular terminal sliding mode control of automotive electronic throttle systems using finite-time exact observer. IEEE Trans. Ind. Electron. 2018, 65, 7160–7172. [Google Scholar] [CrossRef]
  196. Yao, X.; Park, J.H.; Dong, H.; Guo, L.; Lin, X.J. Robust adaptive nonsingular terminal sliding mode control for automatic train operation. IEEE Trans. Syst. ManCybern. Syst. 2018, 49, 2406–2415. [Google Scholar] [CrossRef]
  197. Farid, Y.; Majd, V.J.; Ehsani-Seresht, A.J. Fractional-order active fault-tolerant force-position controller design for the legged robots using saturated actuator with unknown bias and gain degradation. Mech. Syst. Process. 2018, 104, 465–486. [Google Scholar] [CrossRef]
  198. Mousavi, Y.; Alfi, A.J. A memetic algorithm applied to trajectory control by tuning of fractional order proportional-integral-derivative controllers. Appl. Soft Comput. 2015, 36, 599–617. [Google Scholar] [CrossRef]
  199. Wei, Y.-Q.; Liu, D.-Y.; Boutat, D.J. Innovative fractional derivative estimation of the pseudo-state for a class of fractional order linear systems. Automatica 2019, 99, 157–166. [Google Scholar] [CrossRef]
  200. Mousavi, Y.; Alfi, A.J. Fractional calculus-based firefly algorithm applied to parameter estimation of chaotic systems. Chaos Solitons Fractals 2018, 114, 202–215. [Google Scholar] [CrossRef]
  201. Sun, G.; Ma, Z.; Yu, J.J. Discrete-time fractional order terminal sliding mode tracking control for linear motor. IEEE Trans. Ind. Electron. 2017, 65, 3386–3394. [Google Scholar] [CrossRef]
  202. Wang, Y.; Chen, J.; Yan, F.; Zhu, K.; Chen, B.J. Adaptive super-twisting fractional-order nonsingular terminal sliding mode control of cable-driven manipulators. ISA Trans. 2019, 86, 163–180. [Google Scholar] [CrossRef] [PubMed]
  203. Sun, G.; Ma, Z.J. Practical tracking control of linear motor with adaptive fractional order terminal sliding mode control. IEEE/ASME Trans. Mechatronics 2017, 22, 2643–2653. [Google Scholar] [CrossRef]
  204. Yashar, M.; Zarei, A.; Jahromi, Z.S. Robust adaptive fractional-order nonsingular terminal sliding mode stabilization of three-axis gimbal platforms. ISA Trans. 2022, 123, 98–109. [Google Scholar]
  205. Delghavi, M.B.; Shoja-Majidabad, S.; Yazdani, A. Fractional-order sliding-mode control of islanded distributed energy resource systems. IEEE Trans. Sustain. Energy 2016, 7, 1482–1491. [Google Scholar] [CrossRef]
  206. Aghababa, M.P. A Lyapunov-based control scheme for robust stabilization of fractional chaotic systems. Nonlinear Dyn. 2014, 78, 2129–2140. [Google Scholar] [CrossRef]
  207. Li, C.; Deng, W.J. Remarks on fractional derivatives. Appl. Math. Comput. 2007, 187, 777–784. [Google Scholar] [CrossRef]
  208. Zhang, F.; Li, C.J. Stability analysis of fractional differential systems with order lying in (1, 2). Adv. Differ. Equ. 2011, 2011, 1–17. [Google Scholar] [CrossRef] [PubMed]
  209. Gao, Z.; Liao, X.J. Improved Oustaloup approximation of fractional-order operators using adaptive chaotic particle swarm optimization. J. Syst. Eng. Electron. 2012, 23, 145–153. [Google Scholar] [CrossRef]
  210. Burton, T.; Sharpe, D.; Jenkins, N.; Bossanyi, E. Wind Energy Handbook; Wiley Online Library: Hoboken, NJ, USA, 2001. [Google Scholar]
  211. Kiruthiga, B.J. Implementation of first order sliding mode control of active and reactive power for DFIG based wind turbine. Int. J. Inf. Futur. Res. 2015, 2, 2487–2497. [Google Scholar]
  212. Lima, F.; Watanabe, E.H.; Rodriguez, P.; Luna, A. A simplified model for wind turbine based on doubly fed induction generator. In Proceedings of the 2011 International Conference on Electrical Machines and Systems, Beijing, China, 20–23 August 2011; pp. 1–6. [Google Scholar]
  213. Mechter, A.; Kemih, K.; Ghanes, M.J. Sliding mode control of a wind turbine with exponential reaching law. Acta Polytech. Hung. 2015, 12, 167–183. [Google Scholar]
  214. Patton, R.J.; Frank, P.M.; Clark, R.N. Issues of Fault Diagnosis for Dynamic Systems; Springer Science & Business Media: Berlin, Germany, 2013. [Google Scholar]
  215. Huerta, H.; Loukianov, A.J. E nergy based sliding mode control of Brushless Double-fed Induction Generator. Int. J. Electr. Power Energy Syst. 2021, 130, 107002. [Google Scholar] [CrossRef]
  216. Yang, B.; Yu, T.; Shu, H.; Zhang, Y.; Chen, J.; Sang, Y.; Jiang, L. Passivity-based sliding-mode control design for optimal power extraction of a PMSG based variable speed wind turbine. Renew. Energy 2018, 119, 577–589. [Google Scholar] [CrossRef]
  217. Djilali, L.; Caballero-Barragán, H.; Osuna-Ibarra, L.; Sanchez, E.N.; Loukianov, A.J. Neural Sliding Mode Control of a DFIG Based Wind Turbine with Measurement Delay. IFAC-PapersOnLine 2018, 51, 456–461. [Google Scholar] [CrossRef]
  218. Xiong, L.; Wang, J.; Mi, X.; Khan, M.W.J. Fractional order sliding mode based direct power control of grid-connected DFIG. IEEE Trans. Power Syst. 2017, 33, 3087–3096. [Google Scholar] [CrossRef]
  219. Dong, H.; Su, M.; Liu, K.; Zou, W.J. Mitigation Strategy of Subsynchronous Oscillation Based on Fractional-Order Sliding Mode Control for VSC-MTDC Systems With DFIG-Based Wind Farm Access. IEEE Access 2020, 8, 209242–209250. [Google Scholar] [CrossRef]
  220. Zhang, B.; Pi, Y.; Luo, Y.J. Fractional order sliding-mode control based on parameters auto-tuning for velocity control of permanent magnet synchronous motor. ISA Trans. 2012, 51, 649–656. [Google Scholar] [CrossRef]
  221. Amin, U.S.; Ullah, M.; Sajjad, M.; Cheikh, F.A.; Hijji, M.; Hijji, A.; Muhammad, K. EADN: An Efficient Deep Learning Model for Anomaly Detection in Videos. Mathematics 2022, 10, 1555. [Google Scholar] [CrossRef]
Figure 1. Electrical and mechanical control schemes used in WECS.
Figure 1. Electrical and mechanical control schemes used in WECS.
Fractalfract 06 00447 g001
Figure 2. The four operating regions of WECS. Reprinted with permission from Ref. [8]. Copyright 2016 Elsevier.
Figure 2. The four operating regions of WECS. Reprinted with permission from Ref. [8]. Copyright 2016 Elsevier.
Fractalfract 06 00447 g002
Figure 3. Induction vs. synchronous generators.
Figure 3. Induction vs. synchronous generators.
Fractalfract 06 00447 g003
Figure 4. Enhancement techniques in sliding mode control. Reprinted with permission from Ref. [2]. Copyright 2021 Elsevier.
Figure 4. Enhancement techniques in sliding mode control. Reprinted with permission from Ref. [2]. Copyright 2021 Elsevier.
Fractalfract 06 00447 g004
Figure 5. First-order SMC operational diagram.
Figure 5. First-order SMC operational diagram.
Fractalfract 06 00447 g005
Figure 6. Super-twisting-based SMC operational diagram.
Figure 6. Super-twisting-based SMC operational diagram.
Fractalfract 06 00447 g006
Figure 7. Adaptive super twisting SMC operational diagram.
Figure 7. Adaptive super twisting SMC operational diagram.
Fractalfract 06 00447 g007
Figure 8. Integral SMC operational diagram.
Figure 8. Integral SMC operational diagram.
Fractalfract 06 00447 g008
Figure 9. FLC-based system presented by [180]. Reprinted with permission from Ref. [180]. Copyright 2015 Elsevier.
Figure 9. FLC-based system presented by [180]. Reprinted with permission from Ref. [180]. Copyright 2015 Elsevier.
Fractalfract 06 00447 g009
Figure 10. Fractional-order SMC operational diagram.
Figure 10. Fractional-order SMC operational diagram.
Fractalfract 06 00447 g010
Figure 11. Fractional-order terminal SMC operational diagram.
Figure 11. Fractional-order terminal SMC operational diagram.
Fractalfract 06 00447 g011
Figure 12. Fractional-order super twisting SMC operational diagram.
Figure 12. Fractional-order super twisting SMC operational diagram.
Fractalfract 06 00447 g012
Figure 13. Membership function of fuzzy logic control−based FOSMC scheme. Reprinted with permission from Ref. [2]. Copyright 2021 Elsevier.
Figure 13. Membership function of fuzzy logic control−based FOSMC scheme. Reprinted with permission from Ref. [2]. Copyright 2021 Elsevier.
Fractalfract 06 00447 g013
Figure 14. Schematic of DFIG-based WECS operating under different levels.
Figure 14. Schematic of DFIG-based WECS operating under different levels.
Fractalfract 06 00447 g014
Figure 15. (a,b) Stable domain of fractional-order system. Reprinted with permission from Ref. [220]. Copyright 2012 Elsevier.
Figure 15. (a,b) Stable domain of fractional-order system. Reprinted with permission from Ref. [220]. Copyright 2012 Elsevier.
Fractalfract 06 00447 g015
Figure 16. Schematic of DFIG−based WECS operating under FOTSMC. Reprinted with permission from Ref. [2]. Copyright 2021 Elsevier.
Figure 16. Schematic of DFIG−based WECS operating under FOTSMC. Reprinted with permission from Ref. [2]. Copyright 2021 Elsevier.
Fractalfract 06 00447 g016
Figure 17. Generator speed, active, reactive power tracking. Reprinted with permission from Ref. [2]. Copyright 2021 Elsevier.
Figure 17. Generator speed, active, reactive power tracking. Reprinted with permission from Ref. [2]. Copyright 2021 Elsevier.
Fractalfract 06 00447 g017
Figure 18. Errors, surfaces, and DC-link voltage tracking. Reprinted with permission from Ref. [2]. Copyright 2021 Elsevier.
Figure 18. Errors, surfaces, and DC-link voltage tracking. Reprinted with permission from Ref. [2]. Copyright 2021 Elsevier.
Fractalfract 06 00447 g018
Figure 19. CPU and execution times for IOSMC and FOSMC.
Figure 19. CPU and execution times for IOSMC and FOSMC.
Fractalfract 06 00447 g019
Table 1. Comparative analysis of various controllers for DFIG-WECS.
Table 1. Comparative analysis of various controllers for DFIG-WECS.
 Ref.   Technique Employed   SC PU ED UBD CE  R DS FTC 
 [107]PI
 [108]Backstepping
 [109]SMC
 [110]SMC
 [111]SMC with ERL
 [112]Adaptive
backstepping
 [101]ADR control
using ESO
 [113]Adaptive neuro-fuzzy
 [114]SMC
 [115]Coordinated HOSMC
 [116]HOSMC
 [117]HOSMC
 [118]Adaptive STSMC
System complexity (SC), Parametric uncertainty (PU), External disturbances (ED), Upper bound (UB), Chattering elimination (CE), Robustness (R), Dynamical stability (DS), Finite time convergence (FTC).
Table 2. First-order SMC comparison in WECS.
Table 2. First-order SMC comparison in WECS.
TechniqueRef.YearErrorsSurfacesGenerator, Control, and Converter-Type HardwareRemarks
First-order SMC [123]2008 e P = P r e f P s S P = P r e f P s Standalone
No
 [88]2013 e P = P r e f ( k ) P s ( k )
e Q = Q r e f ( k ) Q s ( k )
S P = P r e f ( k ) P s ( k )
S Q = Q r e f ( k ) Q s ( k )
DFIG
BTB converter
RSC control
No
An SMC in discrete domain is presented to control the DFIG plant using power control. Harmonic spectra are also presented in the paper.
 [124]2016 e d c = I r d c I r d c - r e f
e i g = e 1 g e 2 g
= I g d I g d - r e f I g q I g q - r e f
S d c = e d c
S i g = e i g
PMSG
Diode rectifier
Boost converter
NPC inverter at GSC
Yes
The conventional reaching law is replaced by a new enhanced reaching law to reduce the chattering. A gain K is used to trade off between chattering and reaching. The  e d c is error between boost converter current and reference, whereas e i g is taken as error between reference and actual grid current.
 [126]2020 e p = P r e f P s
e Q = Q r e f Q s
S ω = P r e f P s
S Q = Q r e f Q s
RSC side control
No
Stability analysis of SMC and STSMC are presented and performance is evaluated.
Table 3. High-order SMC schemes comparison for WECS.
Table 3. High-order SMC schemes comparison for WECS.
TechniqueRef.YearErrorsSurfacesGenerator, Control, and Converter-Type HardwareRemarks
High-order SMC [138]2008 S 1 = K o p t Ω e 2 3 2 P 2 ϕ s r i q
S 2 = P r e f ( t ) 3 2 ϕ s r i q ω e
PMSG
Rectifier and inverter
No
Rectifier is controlled using STSMC with two different surfaces for two operation modes, which are optimum power conversion and power regulation.
 [116]2012 e Ω = Ω r e f Ω r
e i = I r d I r d - r e f
S ω = Ω r e f Ω r
S i = I r d I r d - r e f
DFIG
BTB converters
Yes
STSMC-based high-order control scheme is proposed. The simulation is carried out in FAST wind turbine simulator. A chattering-free and robust behavior with less mechanical stress is experienced.
 [139]2012 e Ω = Ω r e f Ω r S ω = 1 J Ω r e f Ω r DFIG
No
A Lyapunov-based modified STSMC is presented with variable gains. The results are validated using extensive simulations.
  [140]  2013 e T = T e , r e f T e
e i = Q s , r e f Q s
S T = T e , r e f 3 p L m V s 2 ω s L s i q r
S i = Q s , r e f + 3 L m V s 2 L s i d r V s ω s L m
DFIG
BTB converters
RSC control
No
STSMC with variable gains is presented in this paper and convergence analysis is provided for DFIG-based WECS.
  [135]  2014 e T = T e , r e f T e
e i = I r d I r d - r e f
S T = T e , r e f T e
S i = I r d I r d - r e f
DFIG
BTB converter
No
An STSMC-based second-order SMC is presented using FAST code
[141]   2016   e Ω = Ω r e f Ω r
e E = E E
   S ω = Ω r e f Ω r
S E = E E
PMSG
IGBT 8837
BTB
Yes
[142]   2017   e i = I r d q I r d q - r e f
e E = I g d q I g d q - r e f
   S i = I r d q I r d q - r e f
S E = I g d q I g d q - r e f
DFIG
BTB converters
RSC and GSC control
No
A fault ride through technique using fuzzy integral terminal STSMC is presented in this paper.
 [143] S T = T r e f + T g
S E = Q g , r e f Q g
DFIG
BTB converters
No
A second-order SMC is proposed and made adaptive using time-varying receding horizon. The conservative bounds are calculated and then adaptation strategy is proposed.
 [144]2019 e φ = φ r - r e f φ r
e T = T e , r e f T e
S φ = φ r - r e f φ r
e T = T e , r e f T e
DFIG
RSC control
No
An STSMC is proposed where the gains of the STSMC are optimally selected using new rooted tree optimization.
 [145]2020 e Ω = Ω r e f Ω r S ω = e ˙ ω + k e ω RSC outer loop control
No
A new structure of SMC based on switching sector is presented to minimize the chattering phenomenon.
 [146]2019 e Ω = Ω r e f Ω r S ω = e ˙ ω + k | e ω | 2 3 s i g n ( e ω ) RSC control
Yes
A new inertial perturb-and-observe MPPT technique is presented. A third-order STSMC is proposed for speed tracking.
 [136]    2020     e E 1 = ( E E )
e E 2 = ( E E ) d t
     S E = λ e E 1 + e E 2 PMSG
Rectifier +
Boost converter +
Inverter
No
An SOSMC is presented for the boost converter in PMSG -based WECS to control the DC link voltage.
[147]   2020 e P ( t ) = P s P s
e Q ( t ) = Q s Q s
S 1 = e P ( t ) + K P e P ( τ ) d τ
S 2 = e Q ( t ) + K Q e Q ( τ ) d τ
DFIG
RSC control
Yes
Am HOSMC is presented with enhancing power quality using six power compensation techniques.
 [126] 2020 e p = P r e f P s
e Q = Q r e f Q s
S ω = P r e f P s
S Q = Q r e f Q s
RSC side control
NO
Stability analysis of SMC and STSMC is presented, and  performance is evaluated.
[137]   2021   e p = P s P r e f
e Q = Q s Q r e f
   S 1 = e P ( t ) + K P e P ( τ ) d τ
S 2 = e Q ( t ) + K Q e Q ( τ ) d τ
DFIG
RSC control
No
Adaptive STSMC is proposed for DFIG-based WECS. The  adaptive gains are given as follows:
w = k 1 s | | S | | 1 2 + r
r = k 2 s | | S | | , where
[ w = w P , w Q ] T , [ r = r P , r Q ] T
Table 4. Integral SMC schemes comparison for WECS.
Table 4. Integral SMC schemes comparison for WECS.
TechniqueRef.YearErrorsSurfacesGenerator, Control, and Converter-Type HardwareRemarks
Integral SMC[153]  2009 e T = T e , r e f + T e +
e i = Q r e f + Q g +
S T = e T + e T d t
S i = e i + e i d t
DFIG
BTB converters
RSC control
Yes
Integral SMC is proposed to remove the pulsations on torque and power at frequency twice the grid frequency. Concept of positive and negative SRF is used.
[110]  2010   e P ( t ) = P s P s
e Q ( t ) = Q s Q s
   S 1 = e P ( t ) + K P e P ( τ ) d τ
S 2 = e Q ( t ) + K Q e Q ( τ ) d τ
DFIG
BTB converters
RSC control
No
DPC of DFIG-based WECS is presented and compared with conventional vector control and look-up-table-based DPC.
[154]  2011   e i = I r d q - r e f + I r d q +    S i = e i + 0 t e i d t DFIG
BTB converters
RSC control
No
Concept of positive and negative SRF is used to remove the pulsations on torque and reactive power using ISMC.
[155] e T = T e , r e f T e
e Q ( t ) = Q s Q s
S 1 = e t ( t ) + K P e t ( τ ) d τ
S 2 = e Q ( t ) + K Q e Q ( τ ) d τ
surface derivative is:
S ˙ = k | S | α N ( S ) s i g n ( S )
RSC control
No
RSC control is presented with a new exponential reaching law introduced in the derivative of surface, where the surface is taken as integral type surface.
[156]  2018   e i = I r d q I r d q - r e f S 1 = e i ( t ) + K 0 t e P ( τ ) d τ +
e i ( t 0 ) + y r ( t ) y r ( t 0 )
PMSG
BTB converters
Yes
An integral SMC with integral and novel proportional integral-based surfaces is proposed.
[157]  2020 e Ω = Ω r e f Ω r
e T = T e , r e f T e
e i = I s d I s d - r e f
S ( x ) = G ( x ( t ) x ( 0 ) )
G 0 t A x ( t ) + B u S D R E ( t ) d t
where x = Ω r , T e , I s d
PMSGIntegral SMC is designed using state dependent Ricatti equation (SDRE) with the nonlinear control law based on Ricatti and Lyapunov equation.
[158]2020 e Ω = Ω r e f Ω r
e i = I r d q I r d q - r e f
S ω = A e ω p L m φ s J L s i q r
S i r d = e i + R r | σ | L r 0 t e i d τ
DFIG
RSC control
No
Comparative analysis of SMC and H control is presented. SMC provided good transient performance whereas H provided less tracking error.
[159]2020 e Ω = Ω r e f Ω r S ( t ) = Ω r + k 1 · Ω r +
k 2 0 t Ω r Ω ropt d τ
PMSG
RSC control
No
The uncertain wind turbine dynamics are controlled and estimated using recurrent neural network. Adaptive ISMC is proposed to track optimum rotation speed.
[160]     2021   e E 1 = ( E E ) d t
e E 2 = ( E E )
e E 2 = ( I L I L )
    S ( e 1 , e 2 , e 3 ) = e 1 , λ + e 2 , Y e 3
with Y = L C
PMSG
Rectifier +
Buck converter +
Inverter
Yes
An SMC technique with new MPPT algorithm is proposed to control the DC link voltage.
 [161] 2021 e P ( t ) = P s P s
e Q ( t ) = Q s Q s
S = C e PMSG
RSC control
Yes
An event-triggered integral SMC is presented with extended state observer. The event triggering is used to minimize the communication between execution units and controller.
Table 5. Terminal-based SMC comparison in WECS.
Table 5. Terminal-based SMC comparison in WECS.
TechniqueRef.YearErrorsSurfacesGenerator, Control, and Converter-Type HardwareRemarks
Terminal SMC [173]2015 e P ( t ) = P s P s
e Q ( t ) = Q s Q s
S p = β d 1 τ 1 + β d 2 τ 2 φ d 1 sign τ 2 + β d 3 τ 2 φ d 2 sign τ 2 S Q = β q 1 τ 3 + β q 2 τ 4 φ q 1 sign τ 4 + β q 3 τ 4 φ q 2 sign τ 4
where β > 0 ,
and  τ 1 = e P , τ 2 = e ˙ P , τ 3 = e Q , τ 4 =   e ˙ 0
DFIG
BTB converters
RSC/GSC control
No
An adaptive fast TSMC is presented in this paper. The GSC is controlled using PI control scheme to mitigate the chattering and remove the control input saturation.
 [174]2018 e P t = P s P s
e Q t = Q s Q s
S 1 = e P + β 0 t e P q / p d τ
S 2 = e Q + β 0 t e Q q / p d τ
DFIG
BTB converters
RSC
No
The transient response of DFIG-WECS is improved using adaptive TSMC. The analysis is conducted for both sub-synchronous and synchronous modes of operation.
 [176]2018 e E = I E α β I E α β S E = e ˙ E + β g α β e E PMSG
Yes
An integral type of terminal SMC is proposed in the paper for the power conversion between grid and GSC. Time-varying gains are adopted for the reduction in control energy waste.
 [175]2019 e ( k ) = E ( k ) E ( k 1 ) S E ( k ) = e ( k ) + β i = 0 k e ( i ) λ PMSG
Rectifier
+ Boost converter
Yes
This paper proposed a discrete time terminal SMC for PMSG WECS. The author achieved 94% efficiency at 9 m/s wind speed.
 [177]2020 e i = I r d q I r d q - r e f
e E = I g d q I g d q
S 1 ( t ) = e ˙ r d + β r d e r d + α d e e r d q / p
S 2 ( t ) = e ˙ r q + β r q e r q + α q e e r q q / p
S E 1 ( t ) = e ˙ g d + β g d e g d + α d g e r d q / p
S E 1 ( t ) = e ˙ g q + β g q e g q + α q g e g q q / p
DFIG
BTB converters
RSC/GSC control
No
A chattering-free ITSMC is proposed where the fuzzy control theory is used to tune the gains.
Table 6. Fuzzy-logic-based SMC comparison in WECS.
Table 6. Fuzzy-logic-based SMC comparison in WECS.
TechniqueRef.YearErrorsSurfacesGenerator, Control, and Converter-Type HardwareRemarks
Fuzzy Logic SMC  [180]  2015 P d c , I d c F L E
e E 1 = E E
e E 2 = e ˙ E 1
e E 3 = E E d t
S = k 1 e E 1 + k 2 e E 2 + k 3 e E 3 PMSG
Diode rectifier
+Boost converter
Yes
The authors proposed fuzzy-logic-based ISMC to track the derived voltage. FL is used to calculate the E . An integral-based SMC is adopted using the FL output.
 [181]2015 P d c , V d c F L I d c , o p t
e E = I d c , o p t I d c
S E = I d c , o p t I d c PMSG
Diode rectifier +
Boost converter
Yes
The authors proposed fuzzy-logic-based SMC to reduce the harmonics in GSC current. FL is used to calculate the I d c , o p t . A double-integral-based SMC is adopted using the FL output.
 [182]2016 e ω = Ω r e f Ω r
e q = q Ω r e f
where q = T e m J B J + d
S Ω = e q + d ^ + c e ω
S q = i d
PMSG
RSC control
No
A fuzzy-based SMC is presented with disturbance observer. The discontinuous part is estimated using fuzzy control theory.
 [183]2019 e i = I r d q I r d q - r e f S i = I r d q I r d q - r e f DFIG
BTB converters
RSC control
No
The authors proposed fuzzy-logic-based SMC to reduce the chattering by replacing the sign function with FL control theory. PSO-GSA optimization is used to adjust the membership functions.
Table 8. Fractional-order control schemes and their variants comparison in WECS.
Table 8. Fractional-order control schemes and their variants comparison in WECS.
TechniqueRef.YearErrorsSurfacesGenerator, Control, and Converter-Type HardwareRemarks
Fractional SMC [102]2016 e Ω = Ω r e f Ω r
e i = I r d q I r d q - r e f
S e ω = e e ω + c D α e e ω
S i = e i + c D α e i
DFIG
RSC control
No
An FOSMC is proposed to mitigate the inherent chattering in SMC without the knowledge of uncertainties boundaries.
 [218]2017 e = e 1 e 2 = P r e f P s Q r e f Q s S = c D α s i g n ( e 1 ) c D α s i g n ( e 2 ) DFIG
BTB converter
RSC DPC
Yes
An FOSMC with an enhanced exponential reaching law is proposed. The FOSMC is used to mitigate the inherent chattering in SMC.
    [219]    2020     e i = I r d q I r d q - r e f      S i = e i + c 5 D α e i DFIG
RSC control connected to
voltage-source-converter
-based multi-terminal
DC system
No
A feedback linearization-based system is used to control the DFIG using FOSMC. The considered wind system is connected to a multi-terminal DC system.
 [103]2020 e i = I s d q I s d q - r e f     e i = I E d q I E d q - r e f S s d = e s d + λ s d D α 1 sig e s d γ 1 S s q = e s q + λ s q D α 1 sig e s q γ 2 S E d = e E d + λ E d D α 1 sig e E d γ 1 S E q = e E q + λ E q D α 1 sig e E q γ 2 PMSG
BTB converters
RSC and GSC
Yes
FOSMC-based control is presented. A  sign function is replace by saturation function.
Table 9. Fractional-order terminal and Fractional-order supertwisting SMC control schemes.
Table 9. Fractional-order terminal and Fractional-order supertwisting SMC control schemes.
TechniqueRef.YearErrorsSurfacesGenerator, Control, and Converter-Type HardwareRemarks
Fractional-order terminal SMC [104]2017 e Ω = Ω r e f Ω r
e i = I r d q I r d q - r e f
e E = E E
S Ω = e Ω + c 5 D α e Ω
S i = e i + c 5 D α e i
S E = e E + c 5 D α e E
DFIG
BTB converters
Yes
An artificial intelligence and fractional-order system are integrated to improve the system efficiency. The  discontinuous terms of controllers are proposed using FL theory. The  experiments are validated using PIL-based workbench.
 [106]2020 e Ω = Ω r e f Ω r
e i = I r d q I r d q - r e f
e E = E E
S Ω = c 1 D α e Ω + c 2 e Ω
S i = c 3 , 5 D α e i + c 4 , 6 e i
S E = c 7 D α e E + c 8 e E
DFIG
BTB converters
Yes
Fractional-order super twisting [105]2017 e Ω = Ω r e f Ω r
e i = I r d q I r d q - r e f
e E = E E
S ω = e Ω + c 5 D α e Ω
S i = e i + c 5 D α e i
S E = e E + c 5 D α e E
DFIG
BTB converters
Yes
A fractional-order terminal control is combined with the attributes of super twisting control. The chattering in fractional-order terminal SMC is improved by using super twisting algorithm
Fractional-order composite SMC [2]2020 e Ω = Ω r e f Ω r
e i = I r d q I r d q - r e f
e E = E E
S Ω = e Ω + c 5 D α e Ω
S i = e i + c 5 D α e i
S E = e E + c 5 D α e E
DFIG
BTB converters
Yes
A sensorless FOTSMC is presented in this paper. The speed is estimated using Gaussian process regression. The sign function is replaced by fuzzy control theory to reduce chattering.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sami, I.; Ullah, S.; Khan, L.; Al-Durra, A.; Ro, J.-S. Integer and Fractional-Order Sliding Mode Control Schemes in Wind Energy Conversion Systems: Comprehensive Review, Comparison, and Technical Insight. Fractal Fract. 2022, 6, 447. https://doi.org/10.3390/fractalfract6080447

AMA Style

Sami I, Ullah S, Khan L, Al-Durra A, Ro J-S. Integer and Fractional-Order Sliding Mode Control Schemes in Wind Energy Conversion Systems: Comprehensive Review, Comparison, and Technical Insight. Fractal and Fractional. 2022; 6(8):447. https://doi.org/10.3390/fractalfract6080447

Chicago/Turabian Style

Sami, Irfan, Shafaat Ullah, Laiq Khan, Ahmed Al-Durra, and Jong-Suk Ro. 2022. "Integer and Fractional-Order Sliding Mode Control Schemes in Wind Energy Conversion Systems: Comprehensive Review, Comparison, and Technical Insight" Fractal and Fractional 6, no. 8: 447. https://doi.org/10.3390/fractalfract6080447

Article Metrics

Back to TopTop