Next Article in Journal
Stability of Vertical Rotations of an Axisymmetric Ellipsoid on a Vibrating Plane
Next Article in Special Issue
RFCT: Multimodal Sensing Enhances Grasping State Detection for Weak-Stiffness Targets
Previous Article in Journal
Large-Signal Stability of the Quadratic Boost Converter Using a Disturbance Observer-Based Sliding-Mode Control
Previous Article in Special Issue
Adaptive Control for an Aircraft Wing System with Hysteresis Nonlinearity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Vibration Suppression of an Input-Constrained Wind Turbine Blade System

College of Information Science and Technology, Zhongkai University of Agriculture and Engineering, Guangzhou 510225, China
*
Author to whom correspondence should be addressed.
Mathematics 2023, 11(18), 3946; https://doi.org/10.3390/math11183946
Submission received: 3 August 2023 / Revised: 11 September 2023 / Accepted: 12 September 2023 / Published: 17 September 2023
(This article belongs to the Special Issue Deep Learning and Adaptive Control, 2nd Edition)

Abstract

:
During the actual wind power generation process, wind turbines are often affected by side effects such as blade vibrations, input constraints, and actuator faults. This can lead to a reduction in power generation efficiency and even result in unforeseen losses. This study discusses a robust adaptive fault-tolerant boundary control approach to address the issues of input-constrained and actuator-fault problems in wind turbine blade vibration control. By employing projection mapping techniques and hyperbolic tangent functions, a novel robust adaptive controller based on online dynamic updates is constructed to constrain vibrations, compensate for unknown disturbance upper bounds, and ensure the robustness of the coupled system. Additionally, considering the possibility of actuator faults during the control process, a fault-tolerant controller is proposed to effectively suppress elastic vibrations in the wind turbine blade system even in the presence of actuator faults. The effectiveness of the proposed controller is validated through numerical simulations.

1. Introduction

In recent years, the wind power industry has experienced rapid growth [1,2,3]. Scholars have been actively investigating the appropriate control of wind turbines to optimize wind energy capture during power generation. The blades play a crucial role in absorbing energy from the wind. However, the presence of aeroelastic flutter in the generator can lead to blade vibrations during rotation [4,5]. Severe flutter poses a significant risk of turbine damage. To address this issue, it is essential to conduct an effective dynamic analysis of wind turbine blades and establish efficient control strategies for suppressing blade flutter. Various methods have been proposed to mitigate blade vibrations, including tuned liquid column dampers, tuned mass dampers, and vibration controllers based on active equipment. Compared to the first two methods, active control offers higher flexibility and robustness. For example, in [6], a novel bandwidth selection methodology for the Vold–Kalman filter was proposed to address the issue of unreliable monitoring in wind turbine systems. A significant scheme was proposed for Vold–Kalman filter bandwidth selection to guarantee the consistency and accuracy of the offshore wind turbine condition monitoring process, ensuring reliable fault diagnosis in [7]. Both of the above-mentioned methods can effectively address the vibration issues in wind turbine systems. However, the modes considered are limited. In order to consider all modes of the system, we model the wind turbine blade as a distributed parameter system (DPS) [8,9,10] in this study and adopt a boundary control (BC) scheme [11,12] to consider all system modes, effectively avoiding issues related to mode truncation and overflow.
Compared to other active control strategies, boundary control (BC) offers the advantage of requiring fewer sensors and actuators during its implementation. As a result, it has drawn significant interest from researchers in recent years. For instance, in [13], BC technology was applied to a tall building incorporating an active vibration control system, effectively dealing with input and output constraints. Similarly, in [14], a single limit controller was utilized for the flexible tail of an autonomous robotic fish to achieve complex oscillating bodily actions. In [15], an innovative adaptive offset approach was proposed for a flexible manipulator to handle unpredicted system dynamics and a countless number of time-varying actuator faults. Moreover, in [16], a unique out-of-sync boundary control design was introduced for Markov jump reaction-diffusion neural networks. Despite these impressive developments, the research on BC methods specifically aimed at suppressing wind turbine blade vibrations is still limited, which serves as the driving force behind the motivation for the present study.
Numerous techniques have been developed to address the issue of actuator input constraints, catering to both linear and nonlinear systems. For instance, in [17], a fresh formulation was introduced to enhance tracking performance while ensuring stability, even when dealing with various constraints and unaccounted dynamics. In [18], a distributed cooperative control method for multiple unmanned aerial vehicles (UAVs) with directed topology was studied. Similarly, in [19], the obstacle of automatic carrier landing of a UAV was investigated, taking into account external disturbance and input constraints. Moreover, the distributed formation tracking control problem with multiple leaders under actuator faults and constraints was addressed in [20]. However, despite these developments, there has been limited exploration in addressing the collective influence of input constraints in flexible wind turbine blade systems. This research gap has motivated our study, driving us to seek a solution to this specific problem.
During practical production and control processes, actuators may encounter unknown faults, such as partial failures or getting stuck. These faults can lead to changes in system parameters, causing the actual actuator output to deviate from the desired set value and adversely affecting the overall control performance. To address these issues, a well-designed fault-tolerant control (FTC) scheme becomes essential. In recent research, various adaptive FTC approaches have been presented for different systems. In [21], an adaptive quantized FTC was proposed for a 2-DOF helicopter system with actuator failures and dead-zone nonlinearity. Additionally, an innovative adaptive fuzzy FTC was proposed for a three-dimensional riser-vessel system under input backlash nonlinearity [22]. These advancements represent significant progress in the field of FTC. However, the control design for a wind turbine blade system with input constraints, actuator faults, and unknown disturbances remains an unresolved issue, which motivates further investigation efforts to develop effective solutions in this area.
This paper aims at suppressing the instability of a wind turbine blade system under the input constraint and actuator faults. The following presents a concise summary of the primary contributions: (i) Compared to reference [23], this paper takes into consideration the distributed disturbances affecting the blades and updates the model accordingly. This update improves the fidelity of the system’s dynamics, making it more in line with the actual real-world model. (ii) A novel approach by utilizing the hyperbolic tangent function (HTF) is introduced for constructing adaptive controllers. Additionally, the controllers are coupled with dynamically updated adaptive laws based on the projection mapping operator (PMO), effectively suppressing vibrations and ensuring the robustness of the system. (iii) This study effectively addresses the inherent vibration problem in wind turbine blades, and the proposed approach ensures that the system control does not exhibit any spillover effect.

2. Problem Statement

2.1. System Description

Figure 1 depicts a wind turbine system subjected to unknown external disturbances. In this paper, we consider the dynamic model of the flexible wind turbine blade whose governing equations are as follows [23]
ς y h c Λ ¨ ( ϰ , t ) ς Ξ ¨ ( ϰ , t ) B R Ξ ( ϰ , t ) Π B R Ξ ˙ ( ϰ , t ) + f b ( ϰ , t ) = 0 ,
T R Λ ( ϰ , t ) + ς y h c Ξ ¨ ( ϰ , t ) + Π T R Λ ˙ ( ϰ , t ) Υ Λ ¨ ( ϰ , t ) ψ f b ( ϰ , t ) = 0 .
The boundary conditions governing the tip vibration as presented:
B R Ξ ( τ , t ) + Π B R Ξ ˙ ( τ , t ) F 1 ( t ) f 1 ( t ) = 0 ,
T R Λ ( τ , t ) + Π T R Λ ˙ ( τ , t ) F 2 ( t ) f 2 ( t ) = 0 ,
Ξ ( 0 , t ) = Ξ ( 0 , t ) = Ξ ( τ , t ) = Λ ( 0 , t ) = 0 ,
where y h c denotes the distance from the blade’s center of mass to its shear center. The aerodynamic center offset from the shear center of the blade is represented by ψ . ς signifies the density of the blade, Υ denotes the axial moment of inertia of the flexible blade cross-section, B R represents the bending rigidity, T R is the torsion rigidity, f b ( ϰ , t ) denotes the distributed disturbance, R denotes the set of real numbers, F 1 ( t ) and F 2 ( t ) symbolize the control inputs, and Π indicates the damping coefficient. Additionally, f 1 ( t ) and f 2 ( t ) are the unidentified disturbances at the blade tip ϰ = τ .
Remark 1. 
To enhance clarity, the following symbol substitutions are used: ( · ) ˙ = ( · ) / t , ( · ) = ( · ) / ϰ , ( · ) ˙ = 2 ( · ) / ϰ t , ( · ) = 2 ( · ) / ϰ 2 , and ( · ) ¨ = 2 ( · ) / t 2 .

2.2. Actuator Fault

This paper considers the form of the actuator faults as described
F j ( t ) = μ j F d j ( t ) + F f j ( t ) ,
where 0 < μ j 1 , j = 1 , 2 , represent the extent of actuator faults, F d j , j = 1 , 2 , denote the ideal inputs of the actuator, and F f j , j = 1 , 2 are bounded and unknown functions, which represent the floating faults. Thus, combining (6), the boundary conditions (3) and (4) can be rewritten as
B R Ξ ( τ , t ) + Π B R Ξ ˙ ( τ , t ) μ 1 F d 1 ( t ) F f 1 ( t ) f 1 ( t ) = 0 ,
T R Λ ( τ , t ) + Π T R Λ ˙ ( τ , t ) μ 2 F d 2 ( t ) F f 2 ( t ) f 2 ( t ) = 0 .
We define the “disturbance-like” terms υ j ( t ) , j = 1 , 2 , including external disturbances f j ( t ) , j = 1 , 2 , and the floating faults F f j ( t ) , j = 1 , 2 .
Remark 2. 
Equation (6) represents a conventional actuator fault expression. The parameters μ j , where j = 1 , 2 , determine the efficiency of the actuators. When μ j = 1 , it explains that the actuators are under normal operation. In contrast, if μ j = 0 , it illustrates the actuators have lost all power. We ponder the scenario where the actuator efficiency is partially lost, i.e., μ j ( 0 , 1 ] , for j = 1 , 2 in this study. The parameters F f j ( t ) , where j = 1 , 2 , are time-varying and constrained within the range [ F ¯ f j , F ¯ f j ] with F ¯ f j > 0 .

2.3. Preliminaries

Assumption 1. 
We make the assumption that there are positive constants Δ 1 , Δ 2 , and f h max such that | υ 1 ( t ) | Δ 1 , | υ 2 ( t ) | Δ 2 , and | f b ( ϰ , t ) | f h max , for all ( ϰ , t ) [ 0 , τ ] × [ 0 , + ) .
Lemma 1 
([24]). By assuming the existence of α 1 ( ϰ , t ) and α 2 ( ϰ , t ) R , and ( ϰ , t ) [ 0 , τ ] × [ 0 , + ) , and considering m > 0 , the following expression can be derived
α 1 α 2 α 1 α 2 α 1 2 + α 2 2 ,
α 1 α 2 1 m α 1 2 + m α 2 2 .
Lemma 2 
([24]). Assuming the existence of α ( ϰ , t ) R , and ( ϰ , t ) [ 0 , τ ] × [ 0 , + ) , which satisfying the condition α ( 0 , t ) = 0 , we can express it as follows
α 2 τ 0 τ α 2 d ϰ ,
where ( ϰ , t ) [ 0 , τ ] × [ 0 , + ) .
Lemma 3 
([24]). For every α ( t ) R , the subsequent inequality holds
0 < α α tanh ( α ) β ,
where β = 0.2785 .

3. Control Design

First, we establish the “disturbance-like” components, encompassing external disturbances and the floating faults. Subsequently, an adaptive FTC strategy is formulated to compensate for the actuator fault, handle the composite disturbance, dampen vibration, and maintain displacements within a predefined range.
The control inputs are designed as
F d 1 ( t ) = ρ 1 p ^ 1 [ ν 1 Ξ ( τ , t ) + ν 2 Ξ ˙ ( τ , t ) ] p ^ 1 tanh [ ν 1 Ξ ( τ , t ) + ν 2 Ξ ˙ ( τ , t ) ] Δ ^ 1 ,
F d 2 ( t ) = ρ 2 p ^ 2 [ ν 1 Λ ( τ , t ) + ν 2 Λ ˙ ( τ , t ) ] p ^ 2 tanh [ ν 1 Λ ( τ , t ) + ν 2 Λ ˙ ( τ , t ) ] Δ ^ 2 ,
where ν 1 , ν 2 , ρ 1 , and ρ 2 are positive numbers. Let p j = 1 μ j , j = 1 , 2 , and p ^ j , j = 1 , 2 , denote the estimated values of p j , while the estimated errors are identified as p ˜ j = p j p ^ j , j = 1 , 2 . Δ ^ j , j = 1 , 2 , denote the estimated values of Δ j , j = 1 , 2 , while the estimated errors are identified as Δ ˜ j = Δ j Δ ^ j , j = 1 , 2 .
Step 1. A novel robust adaptive control using PMO and HTF is employed to guarantee the stabilization of the wind turbine blade system.
Δ ^ ˙ 1 ( t ) = γ 1 [ ν 1 Ξ ( τ , t ) + ν 2 Ξ ˙ ( l , t ) ] tanh [ ν 1 Ξ ( τ , t ) + ν 2 Ξ ˙ ( τ , t ) ] γ 1 Δ ^ 1 ( t ) ,
Δ ^ ˙ 2 ( t ) = γ 2 [ ν 1 Λ ( τ , t ) + ν 2 Λ ˙ ( τ , t ) ] tanh [ ν 1 Λ ( τ , t ) + ν 2 Λ ˙ ( τ , t ) ] γ 2 Δ ^ 2 ( t ) ,
p ^ ˙ 1 ( t ) = Proj p ^ 1 { γ 3 [ ν 1 Ξ ( τ , t ) + ν 2 Ξ ˙ ( τ , t ) ] × { ρ 1 tanh [ ν 1 Ξ ( τ , t ) + ν 2 Ξ ˙ ( τ , t ) ] } } p ^ 1 ,
p ^ ˙ 2 ( t ) = Proj p ^ 2 { γ 4 [ ν 1 Λ ( τ , t ) + ν 2 Λ ˙ ( τ , t ) ] × { ρ 2 tanh [ ν 1 Λ ( τ , t ) + ν 2 Λ ˙ ( τ , t ) ] } } p ^ 2 ,
where γ i , i = 1 , , 4 are positive numbers, and Proj p ^ j , j = 1 , 2 , denote the PMO, which restricts the parameters p ^ j , j = 1 , 2 , within the interval [ p ̲ j , p ¯ j ] , j = 1 , 2 [22].
Step 2. A Lyapunov function Ω ( t ) is selected as
Ω ( t ) = Ω 1 ( t ) + Ω 2 ( t ) + Ω 3 ( t ) ,
where
Ω 1 ( t ) = ν 2 2 ς 0 τ [ Ξ ˙ ( ϰ , t ) ] 2 d ϰ + ν 2 2 B R 0 τ [ Ξ ( ϰ , t ) ] 2 d ϰ + ν 2 2 Υ 0 τ [ Λ ˙ ( ϰ , t ) ] 2 d ϰ + ν 2 2 T R 0 τ [ Λ ( ϰ , t ) ] 2 d ϰ ,
Ω 2 ( t ) = ν 1 ς 0 τ Ξ ˙ ( ϰ , t ) Ξ ( ϰ , t ) d ϰ + ν 1 Υ 0 τ Λ ˙ ( ϰ , t ) Λ ( ϰ , t ) d ϰ ν 1 ς y h c 0 τ [ Λ ( ϰ , t ) Ξ ˙ ( ϰ , t ) + Λ ˙ ( ϰ , t ) Ξ ( ϰ , t ) ] d ϰ ν 2 ς y h c 0 τ Ξ ˙ ( ϰ , t ) Λ ˙ ( ϰ , t ) d ϰ ,
Ω 3 ( t ) = 1 2 γ 1 Δ ˜ 1 2 + 1 2 γ 2 Δ ˜ 2 2 + μ 1 2 γ 3 p ˜ 1 2 + μ 2 2 γ 4 p ˜ 2 2 .
Remark 3. 
In this context, Ω 1 ( t ) is derived from the system’s kinetic energy E k ( t ) and potential energy E p ( t ) and is also referred to as the energy term. The term Ω 2 ( t ) arises from the coupling of various state variables within the system, manifesting as a cross-term. Additionally, Ω 3 ( t ) is associated with auxiliary states, denoted as the auxiliary term, representing auxiliary components within the system designed to address disturbance errors.
Theorem 1. 
The function (19) is verified to be positive
0 o 2 [ ζ ( t ) + Ω 3 ( t ) ] Ω ( t ) o 1 [ ζ ( t ) + Ω 3 ( t ) ] ,
where o 1 , and o 2 > 0 .
Proof. 
An auxiliary function is presented as
ζ ( t ) = 0 τ [ Ξ ˙ ( ϰ , t ) ] 2 + [ Ξ ( ϰ , t ) ] 2 + [ Λ ˙ ( ϰ , t ) ] 2 + [ Λ ( ϰ , t ) ] 2 d ϰ .
Thus, we have
ϱ 2 ζ ( t ) Ω 1 ( t ) ϱ 1 ζ ( t ) ,
where ϱ 1 , ϱ 2 > 0 , and ϱ 1 = ν 2 2 max T R , Υ , ς , B R , and ϱ 2 = ν 2 2 min T R , B R , ς , Υ . For Ω 2 ( t ) , we can arrive at
| Ω 2 ( t ) | ν 1 ς 0 τ [ Ξ ˙ ( ϰ , t ) ] 2 d ϰ + τ 4 0 τ [ Ξ ( ϰ , t ) ] 2 d ϰ + ν 1 Υ 0 τ [ Λ ˙ ( ϰ , t ) ] 2 d ϰ + τ 2 0 τ [ Λ ( ϰ , t ) ] 2 d ϰ + ν 1 ς y h c 0 τ [ Ξ ˙ ( ϰ , t ) ] 2 d ϰ + 0 τ [ Λ ˙ ( ϰ , t ) ] 2 d ϰ + ν 1 ς y h c τ 4 0 τ [ Ξ ( ϰ , t ) ] 2 d ϰ + τ 2 0 τ [ Λ ( ϰ , t ) ] 2 d ϰ + ν 2 ς y h c 0 τ [ Ξ ˙ ( ϰ , t ) ] 2 d ϰ + 0 τ [ Λ ˙ ( ϰ , t ) ] 2 d ϰ = ( ν 1 ς + ν 1 ς y h c + ν 2 ς y h c ) 0 τ [ Ξ ˙ ( ϰ , t ) ] 2 d ϰ + ( ν 1 ς + ν 1 ς y h c ) τ 4 0 τ [ Ξ ( ϰ , t ) ] 2 d ϰ + ( ν 1 Υ + ν 1 ς y h c + ν 2 ς y h c ) 0 τ [ Λ ˙ ( ϰ , t ) ] 2 d ϰ + ( ν 1 Υ + ν 1 ς y h c ) τ 2 0 τ [ Λ ( ϰ , t ) ] 2 d ϰ ϱ 3 ζ ( t ) ,
where ϱ 3 = max { ν 1 ς + ν 1 ς y h c + ν 2 ς y h c , ( ν 1 ς + ν 1 ς y h c ) τ 4 , ν 1 Υ + ν 1 ς y h c + ν 2 ς y h c , ( ν 1 Υ + ν 1 ς y h c ) τ 2 } and ν 2 > 2 ϱ 3 min B R , Υ , T R , ς .
Then, (23) can be proved:
0 o 2 [ ζ ( t ) + Ω 3 ( t ) ] Ω ( t ) o 1 [ ζ ( t ) + Ω 3 ( t ) ] ,
where o 2 = ϱ 2 ϱ 3 and o 1 = ϱ 1 + ϱ 3 .
Step 3. Moreover, the Ω ˙ 1 ( t ) is given as
Ω ˙ 1 ( t ) ν 2 Ξ ˙ ( τ , t ) [ B R Ξ ( τ , t ) + Π B R Ξ ˙ ( τ , t ) ] + ν 2 Λ ˙ ( τ , t ) [ T R Λ ( τ , t ) + Π T R Λ ˙ ( τ , t ) ] + ν 2 ς y h c 0 τ [ Λ ¨ ( ϰ , t ) Ξ ˙ ( ϰ , t ) + Λ ˙ ( ϰ , t ) Ξ ¨ ( ϰ , t ) ] d ϰ ν 2 Π T R 2 0 τ [ Λ ˙ ( ϰ , t ) ] 2 d ϰ ( ν 2 Π T R 2 τ 2 η 2 ν 2 ψ ) 0 τ [ Λ ˙ ( ϰ , t ) ] 2 d ϰ ( ν 2 Π B R 2 τ 4 η 1 ν 2 ) 0 τ [ Ξ ˙ ( ϰ , t ) ] 2 d ϰ + ( ν 2 η 1 + ν 2 ψ η 2 ) τ f h m a x 2 ν 2 Π B R 2 0 τ [ Ξ ˙ ( ϰ , t ) ] 2 d ϰ .
Next, we differentiate Ω 2 ( t ) and get
Ω ˙ 2 ( t ) = ν 1 Υ 0 τ Λ ¨ ( ϰ , t ) Λ ( ϰ , t ) d ϰ + ν 1 ς 0 τ Ξ ¨ ( ϰ , t ) Ξ ( ϰ , t ) d ϰ ν 2 ς y h c 0 τ Λ ˙ ( ϰ , t ) Ξ ¨ ( ϰ , t ) d ϰ ν 1 ς y h c 0 τ Λ ( ϰ , t ) Ξ ¨ ( ϰ , t ) d ϰ
ν 1 ς y h c 0 τ Ξ ( ϰ , t ) Λ ¨ ( ϰ , t ) d ϰ + ν 1 ς 0 τ [ Ξ ˙ ( ϰ , t ) ] 2 d ϰ + ν 1 Υ 0 τ [ Λ ˙ ( ϰ , t ) ] 2 d ϰ 2 ν 1 ς y h c 0 τ [ Λ ( ϰ , t ) Ξ ˙ ( ϰ , t ) ] ˙ d ϰ ν 2 ς y h c 0 τ Λ ¨ ( ϰ , t ) Ξ ˙ ( ϰ , t ) d ϰ .
Similarly, we can obtain the Ω ˙ 2 ( t )
Ω ˙ 3 ( t ) = 1 γ 1 Δ ˜ 1 Δ ^ ˙ 1 1 γ 2 Δ ˜ 2 Δ ^ ˙ 2 μ 1 γ 3 p ˜ 1 p ^ ˙ 1 μ 2 γ 4 p ˜ 2 p ^ ˙ 2 .
Considering p ˜ j ( ) p ˜ j Proj p ^ j ( ) 0 , j = 1 , 2 , we derive
Ω ˙ ( t ) ρ 2 [ ν 1 Λ ( τ , t ) + ν 2 Λ ˙ ( τ , t ) ] 2 ρ 1 [ ν 1 Ξ ( τ , t ) + ν 2 Ξ ˙ ( τ , t ) ] 2 ( ν 1 B R ν 1 Π B R η 3 η 6 ν 1 τ 4 ) 0 τ [ Ξ ( ϰ , t ) ] d ϰ ( ν 1 T R ν 1 Π T R η 4 η 7 τ 2 ψ ) 0 τ [ Λ ( ϰ , t ) ] 2 d ϰ ( ν 2 Π B R 2 ν 1 Π B R η 3 ) 0 τ [ Ξ ˙ ( ϰ , t ) ] 2 d ϰ ( ν 2 Π T R 2 ν 1 Π T R η 4 ) 0 τ [ Λ ˙ ( ϰ , t ) ] 2 d ϰ ( ν 2 Π B R 2 τ 4 η 1 ν 2 ν 1 ς 2 ν 1 ς y h c η 5 ) 0 τ [ Ξ ˙ ( ϰ , t ) ] 2 d ϰ γ 1 2 Δ ˜ 1 2 γ 2 2 Δ ˜ 2 2 μ 2 2 γ 4 p ˜ 2 2 ( ν 2 Π T R 2 l 2 η 2 ν 2 ψ ν 1 Υ 2 ν 1 ς y h c η 5 ) 0 τ [ Λ ˙ ( ϰ , t ) ] 2 d ϰ + γ 1 2 Δ 1 2 + γ 2 2 Δ 2 2 μ 1 2 γ 3 p ˜ 1 2 + ( ν 2 η 1 + ν 2 ψ η 2 + ν 1 η 6 + ν 1 ψ η 7 ) τ f h m a x 2 + m ( Δ 1 + Δ 2 ) + μ 1 2 γ 3 p 1 2 + μ 2 2 γ 4 p 2 2 o 3 ζ ( t ) + ϖ ,
where η i > 0 , i = 1 7 .
The intermediate parameters are chosen to satisfy
θ 1 = ν 2 Π B R 2 τ 4 η 1 ν 2 ν 1 ς 2 ν 1 ς y h c η 5 > 0 ,
θ 2 = ν 2 Π T R 2 τ 2 η 2 ν 2 ψ ν 1 Υ 2 ν 1 ς y h c η 5 > 0 ,
θ 3 = ν 1 B R ν 1 Π B R η 3 η 6 ν 1 τ 4 > 0 ,
θ 4 = ν 1 T R ν 1 Π T R η 4 η 7 ν 1 τ 2 ψ > 0 ,
ϖ = ϖ 0 τ f h m a x 2 + γ 2 2 Δ 2 2 + γ 1 2 Δ 1 2 + μ 1 2 γ 3 p 1 2 + μ 2 2 γ 4 p 2 2 + m ( Δ 1 + Δ 2 ) < + ,
where ϖ 0 = ( ν 2 η 1 + ν 2 ψ η 2 + ν 1 η 6 + ν 1 ψ η 7 ) .
Therefore, we can obtain
Ω ˙ ( t ) o Ω ( t ) + ϖ ,
where o = o 3 / o 1 , o 3 = min θ 1 , θ 2 , θ 3 , θ 4 > 0 . □
Theorem 2. 
For the wind turbine blade system represented by (1)–(5) under the influence of the implemented adaptive controllers (13) and (14) and adaptive updating rules (15)–(18), assuming the designed parameters ν 1 , ν 2 , ρ 1 , ρ 2 , and γ i , for i = 1 , , 4 , meeting the constraint conditions (32)–(36) and the initial conditions are bounded.
Proof. 
By using Equation (37) and multiplying both sides by e ϑ t , we obtain
Ω ˙ ( t ) e ϑ t ϑ Ω ( t ) e ϑ t + ϖ e ϑ t .
By integrating Equation (38), we get
Ω ( t ) [ Ω ( 0 ) ϖ ϑ ] e ϑ t + ϖ ϑ Ω ( 0 ) e ϑ t + ϖ ϑ .
Furthermore, by utilizing Equation (23) and applying Lemma 2, we obtain
1 τ 3 [ Ξ ( ϰ , t ) ] 2 1 τ 2 0 τ [ Ξ ( ϰ , t ) ] 2 d ϰ 0 τ [ Ξ ( ϰ , t ) ] 2 d ϰ ζ ( t ) 1 ϑ Ω ( t ) .
Finally, we obtain
Ξ ( ϰ , t ) τ 3 o 2 ( Ω ( 0 ) e ϑ t + ϖ ϑ ) , ( ϰ , t ) [ 0 , τ ] × [ 0 , + ) ,
Λ ( ϰ , t ) τ o 2 ( Ω ( 0 ) e ϑ t + ϖ ϑ ) , ( ϰ , t ) [ 0 , τ ] × [ 0 , + ) .

4. Simulation

Some numerical methods such as the finite difference method, the finite element method, and the finite volume method can be applied to get the approximate solution of the system (1)–(5) when there is no obtainable analytical solution. Among these methods, the finite difference method is relatively simple and well-established, and its advantages are as described in Remark 4. Therefore, in this paper, the finite difference method is chosen for conducting simulation experiments. The system parameters as listed below: B R = 3 Nm 2 , ψ = 0.05 N , ς = 4 kg / m , τ = 2.0 m , y h c = 0.25 m , Υ = 10 kgm , Π = 0.6 , and T R = 20 Nm 2 . The initial states of the system are set to Ξ ( ϰ , 0 ) = ϰ / τ and Λ ( ϰ , 0 ) = π ϰ 2 τ , along with Ξ ˙ ( ϰ , 0 ) = Λ ˙ ( ϰ , 0 ) = 0 . The external disturbances of the system are detailed as follows [23].
f 1 ( t ) = 0.02 + 0.06 s i n ( 0.05 t ) f 2 ( t ) = 0.04 + 0.02 s i n ( 0.1 t )
For simulating the actuator efficacy fault, we configure the efficiency factors as μ 1 = μ 2 = 0.5 at t = 2.5 s , and F f 1 ( t ) = F f 2 ( t ) = sin ( 0.1 t ) .
Figure 2 and Figure 3 depict the deformation of the flexible wind turbine blade without control, i.e., when F 1 ( t ) = 0 and F 2 ( t ) = 0 . Figure 4 and Figure 5 illustrate the stereoscopic representation of the flexible wind turbine blade with the proposed control. The controllers are parameterized with ρ 1 = 2 , ρ 2 = 1 , γ 1 = 0.02 , γ 2 = 5 , γ 3 = 0.1 , and γ 4 = 3 . The input signal constraints in this paper are set as F ¯ 1 = 3 and F ¯ 2 = 3.5 . We conducted simulations for the following two cases: (1) In case 1, we adopted the control scheme proposed as (44). (2) In case 2, we used the control laws (13)–(18) proposed in this paper for the simulation.
F d 1 ( t ) = ρ 1 [ ν 1 Ξ ( τ , t ) + ν 2 Ξ ˙ ( τ , t ) ] F d 2 ( t ) = ρ 2 [ ν 1 Λ ( τ , t ) + ν 2 Λ ˙ ( τ , t ) ]
When employing the control scheme proposed by [25], it becomes evident from Figure 6 and Figure 7 that the control inputs required to achieve the same control objectives exceed the input constraints significantly. The actual inputs’ maximum value even reaches six times the input upper limit, which poses serious risks in practical engineering scenarios with input signal constraints. In contrast, the controllers proposed in this paper with input constraints, as shown in Figure 8 and Figure 9, ensure that the tip forces F 1 and F 2 strictly remain within the specified constraints.
As observed in Figure 8 and Figure 9, there is a noticeable jump in the control input at 2.5 s , which corresponds to the moment of the actuator fault. However, despite the actuator fault, the preset control targets can still be achieved, demonstrating the strong fault tolerance capability of the loop system.
Remark 4. 
The finite difference method is an approximate numerical solution that directly turns a differential problem into an algebraic problem. The mathematical concept is intuitive and the expression is simple. It is an early and mature numerical method. The finite difference method can solve many kinds of partial differential equations systems easily [26]. So in this manuscript, we use the finite difference method to finish the simulations.

5. Conclusions

In this study, an innovative and robust adaptive FTC strategy is proposed for a wind turbine blade system with constrained inputs, which is subjected to uncertainties in the upper bounds of external disturbances. The proposed adaptive FTC approach is specifically designed to attenuate vibrations and rectify actuator faults effectively. By incorporating PMO and HTF into the control method, the system’s robustness is assured, and the inputs are strictly maintained within the imposed constraints. Through rigorous Lyapunov analysis, the controlled system’s uniformly bounded stability is ensured. The effectiveness of the control scheme is validated via comprehensive simulations, where the results are compared to assess the control performance. In the future, we intend to develop a neural network control or reinforcement learning control strategy to handle system uncertainties and external disturbances [27,28].

Author Contributions

Conceptualization, L.C. and S.L.; methodology, L.C.; software, S.L.; validation, L.C. and S.L.; formal analysis, L.C.; investigation, S.L.; resources, L.C. and S.L.; data curation, L.C.; writing—original draft preparation, L.C. and S.L.; supervision, L.C.; project administration, S.L.; funding acquisition, L.C. and S.L. All authors have read and agreed to the published version of the manuscript.

Funding

This paper was supported partly by Guangdong Provincial Graduate Education Innovation Plan Project under Grant 2022XSLT056, Guangdong Provincial Education Science Planning Project under Grant 2020GXJK102, Haizhu District Science and Technology Project under Grant Haike Industrial and Commercial Information Plan 2022-36, Guangdong Agricultural Product Big Data Operation and Application—Construction of Big Data Platform Project under Grant M4400000707018082, the natural Science Foundation of Guangdong Province under Grant2022B1515120059, Innovation Team Project of Universities in Guangdong Province under Grant 2021KCXTD019, Science and Technology Planning Project of Yunfu under Grant 2023020202, 2023020203, 2023020205, Natural Science Foundation of Guangdong Province under Grant 2023A1515011230, Guangdong Provincial College students innovation and entrepreneurship training Plan project under Grant S202211347083, National College students innovation and entrepreneurship training Plan project under Grant 202111347004.

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. Zine, M.; Chemsa, A.; Labiod, C.; Ikhlef, M.; Srairi, K.; Benbouzid, M. Coupled indirect torque control and maximum power point tracking technique for optimal performance of 12/8 switched reluctance generator-based wind turbines. Machines 2022, 10, 1046. [Google Scholar] [CrossRef]
  2. Zweiffel, M.; Jacobs, G.; Song, Z.; Weidinger, P.; Decker, T.; Röder, J.; Bosse, D. Influence of drivetrain efficiency determination on the torque control of wind turbines. Forsch. im Ingenieurwesen 2023, 87, 129–138. [Google Scholar] [CrossRef]
  3. Wang, F.; Chen, J.; Cheng, M.; Xu, B. A novel hydraulic transmission solution to large offshore wind turbine: Design and control strategy. Ocean Eng. 2022, 255, 111285. [Google Scholar] [CrossRef]
  4. Pustina, L.; Serafini, J.; Pasquali, C.; Solero, L.; Lidozzi, A.; Gennaretti, M. A novel resonant controller for sea-induced rotor blade vibratory loads reduction on floating offshore wind turbines. Renew. Sustain. Energy Rev. 2023, 173, 113073. [Google Scholar] [CrossRef]
  5. Zhan, J.; Wu, C.; Ma, X.; Yang, C.; Miao, Q.; Wang, S. Abnormal vibration detection of wind turbine based on temporal convolution network and multivariate coefficient of variation. Mech. Syst. Signal Process. 2022, 174, 109082. [Google Scholar] [CrossRef]
  6. Feng, K.; Ji, J.C.; Ni, Q. A novel adaptive bandwidth selection method for Vold–Kalman filtering and its application in wind turbine planetary gearbox diagnostics. Struct. Health Monit. 2023, 22, 1027–1048. [Google Scholar] [CrossRef]
  7. Feng, K.; Ji, J.C.; Wang, K.; Wei, D.; Zhou, C.; Ni, Q. A novel order spectrum-based Vold-Kalman filter bandwidth selection scheme for fault diagnosis of gearbox in offshore wind turbines. Ocean Eng. 2022, 266, 112920. [Google Scholar] [CrossRef]
  8. Zhang, Q.; Song, X.; Song, S.; Stojanovic, V. Finite-Time sliding mode control for singularly perturbed PDE systems. J. Frankl. Inst. 2023, 360, 841–861. [Google Scholar] [CrossRef]
  9. Zhao, Z.; He, X.; Ahn, C.K. Boundary disturbance observer-based control of a vibrating single-link flexible manipulator. IEEE Trans. Syst. Man Cybern. Syst. 2019, 51, 2382–2390. [Google Scholar] [CrossRef]
  10. Li, R. Boundary exponential stabilization of a one-dimensional anti-stable wave equation with control matched disturbance. Engineering 2020, 12, 640–651. [Google Scholar] [CrossRef]
  11. Zhao, Z.; Liu, Z.; He, W.; Hong, K.S.; Li, H.X. Boundary adaptive fault-tolerant control for a flexible Timoshenko arm with backlash-like hysteresis. Automatica 2021, 130, 109690. [Google Scholar] [CrossRef]
  12. Zhao, Z.; Liu, Y.; Zou, T.; Hong, K.S.; Li, H.X. Robust adaptive fault-tolerant control for a riser-vessel system with input hysteresis and time-varying output constraints. IEEE Trans. Cybern. 2022, 53, 3939–3950. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, S.; Li, Q.; Zhao, X.; Liu, Z.; Li, G. Vibration Control for an Active Mass Damper of a High-Rise Building with Input and Output Constraints. IEEE/ASME Trans. Mechatron. 2022, 28, 186–196. [Google Scholar] [CrossRef]
  14. Zhang, S.; Qian, X.; Liu, Z.; Li, Q.; Li, G. PDE modeling and tracking control for the flexible tail of an autonomous robotic fish. IEEE Trans. Syst. Man Cybern. Syst. 2022, 52, 7618–7627. [Google Scholar] [CrossRef]
  15. Ma, Y.; He, X.; Zhang, S.; Sun, Y.; Fu, Q. Adaptive compensation for infinite number of actuator faults and time-varying delay of a flexible manipulator system. IEEE Trans. Ind. Electron. 2022, 69, 13141–13150. [Google Scholar] [CrossRef]
  16. Han, X.X.; Wu, K.N.; Niu, Y. Asynchronous boundary control of Markov jump neural networks with diffusion terms. IEEE Trans. Cybern. 2022, 53, 4962–4971. [Google Scholar] [CrossRef] [PubMed]
  17. Azizkhani, M.; Zareinejad, M.; Khosravi, M.A. Model reference adaptive control of a soft bending actuator with input constraints and parametric uncertainties. Mechatronics 2022, 84, 102800. [Google Scholar] [CrossRef]
  18. Duan, H.; Yuan, Y.; Zeng, Z. Distributed cooperative control of multiple UAVs in the presence of actuator faults and input constraints. IEEE Trans. Circuits Syst. II Express Briefs 2022, 69, 4463–4467. [Google Scholar] [CrossRef]
  19. Yuan, Y.; Duan, H.; Zeng, Z. Automatic Carrier Landing Control with External Disturbance and Input Constraint. IEEE Trans. Aerosp. Electron. Syst. 2022, 59, 1426–1438. [Google Scholar] [CrossRef]
  20. Jamali, M.; Baghaee, H.R.; Sadabadi, M.S.; Gharehpetian, G.B.; Anvari-Moghaddam, A.; Blaabjerg, F. Distributed finite-time fault-tolerant control of isolated ac microgrids considering input constraints. IEEE Trans. Smart Grid 2022, 13, 4525–4537. [Google Scholar] [CrossRef]
  21. Zhao, Z.; Zhang, J.; Liu, Z.; He, W.; Hong, K.S. Adaptive quantized fault-tolerant control of a 2-DOF helicopter system with actuator fault and unknown dead zone. Automatica 2023, 148, 110792. [Google Scholar] [CrossRef]
  22. Zhao, Z.; Liu, Y.; Ma, G.; Hong, K.S.; Li, H.X. Adaptive Fuzzy Fault-Tolerant Control for a Riser-Vessel System with Unknown Backlash. IEEE Trans. Syst. Man Cybern. Syst. 2023; in press. [Google Scholar] [CrossRef]
  23. Gao, S.; Liu, J. Adaptive fault-tolerant vibration control of a wind turbine blade with actuator stuck. Int. J. Control 2020, 93, 713–724. [Google Scholar] [CrossRef]
  24. Liu, S.; Liu, Z.; Li, Y.; He, W. Nonlinear disturbance observer-based direct joint control for manipulation of a flexible payload with output constraints. Int. J. Control 2023, 96, 1377–1388. [Google Scholar] [CrossRef]
  25. He, W.; Zhang, S. Control design for nonlinear flexible wings of a robotic aircraft. IEEE Trans. Control Syst. Technol. 2016, 25, 351–357. [Google Scholar] [CrossRef]
  26. Fadlun, E.A.; Verzicco, R.; Orlandi, P.; Mohd-Yusof, J. Combined immersed-boundary finite-difference methods for three-dimensional complex flow simulations. J. Comput. Phys. 2000, 161, 35–60. [Google Scholar] [CrossRef]
  27. Zhao, Z.; Ren, Y.; Mu, C.; Zou, T.; Hong, K.S. Adaptive neural-network-based fault-tolerant control for a flexible string with composite disturbance observer and input constraints. IEEE Trans. Cybern. 2021, 52, 12843–12853. [Google Scholar] [CrossRef]
  28. Wu, C.; Pan, W.; Staa, R.; Liu, J.; Sun, G.; Wu, L. Deep reinforcement learning control approach to mitigating actuator attacks. Automatica 2023, 152, 110999. [Google Scholar] [CrossRef]
Figure 1. The wind turbine system.
Figure 1. The wind turbine system.
Mathematics 11 03946 g001
Figure 2. Bending deformations without control.
Figure 2. Bending deformations without control.
Mathematics 11 03946 g002
Figure 3. Twist deformation without control.
Figure 3. Twist deformation without control.
Mathematics 11 03946 g003
Figure 4. Bending deformation with control.
Figure 4. Bending deformation with control.
Mathematics 11 03946 g004
Figure 5. Twist deformation with control.
Figure 5. Twist deformation with control.
Mathematics 11 03946 g005
Figure 6. Actual input F 1 ( t ) in case 1.
Figure 6. Actual input F 1 ( t ) in case 1.
Mathematics 11 03946 g006
Figure 7. Actual input F 2 ( t ) in case 1.
Figure 7. Actual input F 2 ( t ) in case 1.
Mathematics 11 03946 g007
Figure 8. Actual input F 1 ( t ) in case 2.
Figure 8. Actual input F 1 ( t ) in case 2.
Mathematics 11 03946 g008
Figure 9. Actual input F 2 ( t ) in case 2.
Figure 9. Actual input F 2 ( t ) in case 2.
Mathematics 11 03946 g009
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cao, L.; Liu, S. Vibration Suppression of an Input-Constrained Wind Turbine Blade System. Mathematics 2023, 11, 3946. https://doi.org/10.3390/math11183946

AMA Style

Cao L, Liu S. Vibration Suppression of an Input-Constrained Wind Turbine Blade System. Mathematics. 2023; 11(18):3946. https://doi.org/10.3390/math11183946

Chicago/Turabian Style

Cao, Liang, and Shuangyin Liu. 2023. "Vibration Suppression of an Input-Constrained Wind Turbine Blade System" Mathematics 11, no. 18: 3946. https://doi.org/10.3390/math11183946

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop