Next Article in Journal
Extracting Fingerprint Features Using Autoencoder Networks for Gender Classification
Next Article in Special Issue
Research on the Diagnosability of a Satellite Attitude Determination System on a Fault Information Manifold
Previous Article in Journal
Shape Optimization of the Streamlined Train Head for Reducing Aerodynamic Resistance and Noise
Previous Article in Special Issue
Fault Diagnosis Method for Aircraft EHA Based on FCNN and MSPSO Hyperparameter Optimization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Predefined-Time Nonsingular Attitude Control for Vertical-Takeoff Horizontal-Landing Reusable Launch Vehicle

School of Astronautics, Harbin Institute of Technology, Harbin 150001, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(19), 10153; https://doi.org/10.3390/app121910153
Submission received: 14 September 2022 / Revised: 3 October 2022 / Accepted: 6 October 2022 / Published: 9 October 2022

Abstract

:
This paper presents a novel predefined-time nonsingular tracking control system for a vertical-takeoff horizontal-landing (VTHL) reusable launch vehicle (RLV) in the face of parameter uncertainties, model couplings and external disturbances. Firstly, this paper proposes a novel predefined-time prescribed performance function (PTPPF) with desired steady-state and transient performance. The convergence time of PTPPF from the transient state to the steady state can be flexibly adjusted by changing one parameter. Moreover, the decreasing rate of PTPPF in the transient phase can also be adjusted by changing one parameter on the premise of not changing the convergence time of PPF to reach steady state. A novel predefined-time terminal sliding mode surface (SMS) is designed to avoid the singularity, and the attitude tracking errors on SMS are predefined-time stable. By utilizing PTPPF and error transformation, this paper designs a novel nonsingular sliding mode controller to guarantee the attitudes of RLV with desired tracking performance. Without using piecewise functions, the phenomenon of singularity can be avoided. The Lyapunov method is used to verify the stability of the controller. Lastly, a numerical simulation is presented to validate the efficiency of the proposed controller.

1. Introduction

As a fast-response, reliable, and low-cost vehicle to access space, the vertical-takeoff horizontal-landing (VTHL) reusable launch vehicle (RLV) has become a research hotspot in recent years [1,2] The shape diagram of the VTHL RLV studied in this paper is shown in Figure 1. However, due to its complex temperature variation and wide reentry flight envelope, RLV is prone to parameter uncertainties and complex aerodynamic characteristics in the reentry phase, which makes controller design a challenging task. Hence, an advanced reentry attitude control scheme with sufficient robustness and fast convergence rate is significant for RLV to achieve a successful mission.
A variety of control methods have been developed to improve the flight quality of reentry vehicles. Song et al. [3] proposed a fractional-order PID control method for hypersonic vehicle attitude control. Ming et al. [4] proposed an active disturbance rejection controller (ADRC) for an air-breathing supersonic vehicle with a new smooth function. Mao et al. [5] proposed an adaptive fuzzy H-infinity controller for RLV under parameter uncertainties and external disturbances. Zhang et al. [6] designed a backstepping technique and combined a nonlinear observer for hypersonic vehicle, which realized stable tracking of attitude. Xu et al. [7] investigated a controller for hypersonic flight vehicle with robust adaptive neural networks. Compared with the above methods, sliding mode control (SMC), which is robust to disturbances and insensitive to parameter uncertainties, has attracted the most interest [8,9].
The peculiarity of SMC is to adjust the attitude tracking errors of RLV toward a predefined sliding surface and then avoids them leaving this surface. As it is known, the traditional SMC merely ensures the asymptotic stability of the control system. Then, a finite-time SMC with solid robustness against internal and external disturbances was proposed by Lu et al. [10] and Hall et al. [11] to obtain finite-time convergence of states. Since the initial condition of the system determines the settling time of finite-time SMC. It may destroy the desired fast convergence performance when the initial values are unknown in practice. Thus, the fixed-time SMC (FxSMC), which has achieved great development in the past several years, can ensure that the system convergence time is independent of the initial conditions [12,13]. Zhang et al. [14] presented a fixed-time nonsingular terminal SMC for RLV. Ju et al. [15] presented an FxSMC to ensure that the vehicle attitudes track the desired command in a fixed time. Liang et al. [16] designed a robust controller for RLV based on a FxSMC. However, the convergence time of the FxSMC is actually obtained after complex calculation rather than known before controller design, which increases the offline workload. Thus, the predefined-time stability has been widely studied in recent years [17]. The upper bound of the settling time is displayed explicitly as an adjustable control parameter [18]. Torres et al. [19] designed a formulation for robust predefined-time stability of second-order system and proved the reliability of the scheme through simulation. Zhang et al. [20] applied a predefined-time SMC to a launch vehicle attitude control. A robust predefined-time SMC was proposed with external disturbances [21]. However, the application of these predefined-time SMC methods is limited by a singular problem, and the studies on nonsingular predefined-time SMC have rarely been reported.
There is insufficient attention paid to the system transient tracking performance, although the above controllers can guarantee good steady-state control performance. Prescribed performance control, which consists of prescribed performance function (PPF) and error transformation, ensures that the attitude tracking errors converge at a specified rate to a preset arbitrarily small set of residuals while guaranteeing that the maximum overshoot value is smaller than the predefined value. More recently, Bu et al. [22] proposed a prescribed performance controller to obtain an expected transient performance for hypersonic vehicle. Zhang et al. [23] designed a controller with prescribed transient performance to solve attitude control problem. Luo et al. [24] proposed a low-complexity prescribed performance method for vehicles under unknown dynamics errors and external disturbances. However, the above PPF cannot freely adjust its decreasing rate in the transient process without changing the time to reach the steady state. This makes it difficult for PPF to meet different task requirements in practical applications.
To overcome the drawbacks mentioned above, a novel nonsingular predefined-time sliding mode control method for VTHL RLV is proposed. The principal innovations are summarized as follows:
  • A novel predefined-time PPF (PTPPF) is proposed. The convergence time in the transient phase of PTPPF to the steady state can be flexibly adjusted by changing one parameter. Moreover, the decreasing rate of PTPPF in transient phase can also be adjusted by changing one parameter on the premise of not changing the convergence time of PPF to reach steady state.
  • A novel predefined-time sliding mode surface (SMS) is proposed. When the attitude tracking errors moving on the SMS, they can converge to the origin within a predefined time, and the value of convergence time can be adjusted by one parameter.
  • Based on PTPPF and the novel SMS, the RLV control system with a novel nonsingular controller is proposed to prevent the attitude tracking errors from exceeding the prescribed performance bounds. Meanwhile, the singularity can be avoided directly without using piecewise continuous functions.
The remainder of this paper is organized as follows. The dynamic model of RLV and preliminaries are introduced in Section 2. The PTPPF and the error transformation are proposed in Section 3. In Section 4, the nonsingular predefined-time SMS and controller are proposed. The simulation results are shown in Section 5, and Section 6 presents the conclusions of this study.

2. Problem Formulation and Preliminary

2.1. Problem Statement

In this paper, the attitude model of RLV in the reentry phase is described as follows [12]:
{ α ˙ = sin σ cos β [ ψ ˙ cos γ θ ˙ sin ψ sin γ + ( φ ˙ + ω E ) ( cos θ cos ψ sin γ sin θ cos γ ) ] cos σ cos β [ γ ˙ θ ˙ cos ψ ( φ ˙ + ω E ) cos θ sin ψ ] p cos α tan β + q r sin α tan β β ˙ = p sin α r cos α + sin σ [ γ ˙ θ ˙ cos ψ + ( φ ˙ + ω E ) cos θ sin ψ ] + cos σ [ ψ ˙ cos γ θ ˙ sin ψ sin γ ( φ ˙ + ω E ) ( cos θ cos ψ sin γ sin θ cos γ ) ] σ ˙ = p cos α cos β q sin β r sin α cos β + α ˙ sin β ψ ˙ sin γ θ ˙ sin ψ cos γ + ( φ ˙ + ω E ) ( cos θ cos ψ cos γ + sin θ sin γ ) p ˙ = J z z M x J x x J z z J x z 2 + J x z M z J x x J z z J x z 2 + ( J x x J y y + J z z ) J x z J x x J z z J x z 2 p q + ( J y y J z z ) J z z J x z 2 J x x J z z J x z 2 q r q ˙ = M y J y y + J x z ( r 2 p 2 ) J y y + ( J z z J x x ) J y y p r r ˙ = J x z M x J x x J z z J x z 2 + J x x M z J x x J z z J x z 2 + ( J x x J y y + J x z ) J x z J x x J z z J x z 2 p q + ( J y y J z z J x x ) J x z J x x J z z J x z 2 q r ,
where θ and φ represent the latitude and the longitude, respectively. q , r , and p are the pitch angular rate, yaw angular rate, and roll angular rate, respectively. β , α , and σ represent the sideslip, attack, and bank angles, respectively. ψ and γ denote the heading and flight path angles, respectively. J i j ( i = x , y , z ; j = x , y , z ) denotes the moment of inertia, M i ( i = x , y , z ) is the external lumped moment, and ω E is the Earth’s rotation velocity.
Equation (1) can be written as
{ Λ ˙ = R ω + f J ω ˙ = Ω J ω + M a i r + M c + Δ d ,
where Λ = [ α , β , σ ] T , and ω = [ p , q , r ] T . Δ d represents the external disturbances and parameter uncertainties. The inertia matrix is J R 3 × 3 , while M c R 3 denotes the control moment. The matrices R and Ω and the aerodynamic moment M a i r are represented as
R = [ cos α tan β 1 sin α tan β sin α 0 cos α cos α cos β sin β sin α cos β ] ,
Ω = [ 0 r q r 0 p q p 0 ] ,
Μ a i r = [ ( m x p L r V + m x β β ) q 0 S r L r ( m y q L r V + m y α α ) q 0 S r L r ( m z r L r V + m z β β ) q 0 S r L r ] ,
where the air density is ρ , and V is the velocity of RLV. m x β , m y α , and m z β are the static stability moment coefficients of RLV; the damping moment coefficients are m x p , m y q , and m z r . q 0 = 0.5 ρ V 2 . The reference area of RLV is denoted as S r , and the reference length of RLV is denoted as L r .
Notation 1. 
For the given vector x = [ x 1 , x 2 , , x n ] T , define sig a x = [ | x 1 | a sign ( x 1 ) , | x 2 | a sign ( x 2 ) , , | x n | a sign ( x n ) ] T and | x | a = [ | x 1 | a ,   | x 2 | a , , | x n | a ] T , where sign ( ) denotes the sign function and a is a real number.
Assumption 1. 
The external disturbances, parameter uncertainties and their variation rates are assumed to be bounded | Δ d | L d , | Δ d ˙ | L d d , where L d and L d d are unknown positive vectors [25].
Denote z 1 = Λ Λ C and z 2 = Λ ˙ Λ ˙ C . Λ C = [ α C , β C , σ C ] T is the guidance command vector. Then, the attitude tracking error model of RLV can be established as follows:
{ z ˙ 1 = z 2 z ˙ 2 = d + ( R ˙ R 1 + R J 1 Ω J R 1 ) z 2 + R J 1 ( M c + M a i r ) + ( R ˙ R 1 + R J 1 Ω J R 1 ) Λ ˙ C Λ ¨ C ,
where d = R ˙ R 1 f + R J 1 Ω J R 1 f + R J 1 Δ d + f ˙ .
Control Objective: In the presence of the external disturbances and parameter uncertainties, this paper aimed to design a novel nonsingular predefined-time controller for RLV. z 1 can reach a predefined small set of residuals within predefined time. z 1 can always be constrained within the range specified by PTPPF. Meanwhile, the control input M c is nonsingular.

2.2. Preliminary

Consider the autonomous dynamical system
x ˙ = f ( x ; φ ) , x ( 0 ) = 0 ,
where x n denotes the state, and f : n n is a nonlinear locally Lebesgue-integrable function. φ l represents the tunable parameters.
Definition 1 ([26]).  
The origin of Equation (7) is said to be globally stable in predefined time if, for any T c + , there exists some φ l such that x ( t ) = 0 holds for t T c regardless of x ( t 0 ) , where · stands for the Euclidean norm.
Definition 2 ([27]). 
A smooth function ρ ( t ) :   + + is called a performance function if both of the following conditions are satisfied:
(1)
ρ ( t ) is decreasing and positive;
(2)
lim t ρ ( t ) = ρ > 0 .
Lemma 1 ([28]). 
For the system in Equation (7), assume that the Lyapunov function V ( x ) : n 0 satisfies
V ˙ ( x ) π a T p ( V 1 a 2 ( x ) + V 1 + a 2 ( x ) ) ,
where T p > 0 and 0 < a < 1 are positive constants. Then, this system is globally stable in predefined time, and T p is the convergence time.
Theorem 1. 
For the system in Equation (7), if the radially unbounded Lyapunov function V ( x ) : n 0 satisfies
V ˙ ( x ) 3 π + 2 3 arctan 1 3 ( 3 + 2 α ) T c ( V 1 / 2 a / 3 ( x ) + V 2 + 2 a / 3 ( x ) ) ,
where 1.5 < a < 1.5 , T c > 0 , then the system is globally stable in predefined time. Meanwhile, the convergence time is T c .
Proof of Theorem 1. 
Please refer to Appendix A. □
Lemma 2. 
For any y i > 0   ( i = 1 ,   2 ,   ,   n ) , there is
i = 1 n y i γ ¯ ( i = 1 n y i ) γ ¯ ,     if   0 < γ ¯ 1           i = 1 n y i γ ¯ n 1 γ ¯ ( i = 1 n y i ) γ ¯ ,     if   γ ¯ > 1 ,
where γ ¯ > 0 [29].

3. PTPPF and Error Transformation

In this section, a PTPPF is designed, and then the inequality constraint is converted to an unconstrained one, laying a foundation for the nonsingular predefined-time attitude controller design.

3.1. Performance Function Design

To ensure that the transient and steady performance of the tracking errors evolves within the reasonable boundaries, the description of the prescribed performance of the z 1 is expressed as
ρ i ( t ) < z 1 i ( t ) < ρ i ( t ) ,
where i = 1 ,   2 ,   3 , and ρ i ( t ) denotes a novel PTPPF, which is described as follows:
ρ i ( t ) = { a 1 + a 2 arctan ( t T f ) + a 3 e k t T f + a 4 e k ( t T f ) 2       ,   0 t T f ρ i ,     t > T f ,
a 1 = ρ i 0 + 3 a 4   ,       a 2 = 4 k e k a 4 a 3 = 4 a 4             ,       a 4 = ρ i ρ i 0 3 π k e k 3 e k ,
where ρ i 0 is the maximum overshoot, and ρ i is the prescribed steady-state boundary. ρ i 0 and ρ i can be flexibly adjusted according to the control requirements. k is a positive constant, 0.5 < k < 3 , which determines the decreasing rate of ρ i when 0 t T f . T f denotes the prescribed terminal time that ρ i reaches ρ i .
Theorem 2. 
The PTPPF ρ i ( t ) satisfies ρ ˙ i ( t ) < 0 when 0 t T f and lim t T f ρ i = lim t T f + ρ i = ρ i . Meanwhile, it has a continuous first-order derivative with respect to time.
Proof of Theorem 2. 
When 0 t T f , the derivative of Equation (12) is
ρ ˙ i = a 2 T f ( 1 + ( t / T f ) 2 ) k a 3 T f e k t T f 2 k a 4 t T f 2 e k ( t T f ) 2       = 4 k e k a 4 T f ( 1 + ( t / T f ) 2 ) + 4 k a 4 T f e k t T f 2 k a 4 t T f 2 e k ( t T f ) 2 ,
ρ ¨ i = 2 a 2 t T f 3 ( 1 + ( t / T f ) 2 ) 2 k 2 a 3 T f 2 e k t T f 2 k a 4 T f 2 e k ( t T f ) 2 + 4 k 2 a 4 t 2 T f 4 e k ( t T f ) 2 .
According to Equation (13), it is obvious that lim t T f ρ i = lim t T f + ρ i and lim t T f ρ ˙ i = lim t T f   ρ ˙ i = 0 . The PTPPF and its first-order derivative are continuous with respect to time.
Considering Equation (13) and the range of k , it can be calculated that a 4 < 0 . To reduce computational cost, the proof of ρ ˙ i < 0 can be converted to y > 0 on [ 0 ,   T f ] , where y is
y = 4 k e k T f ( 1 + ( t / T f ) 2 ) + 4 k T f e k t T f 2 k t T f 2 e k ( t T f ) 2   ,     0 t < T f .
Let x = t / T f ; Equation (16) can then be simplified as follows:
y = 4 k e k T f ( 1 + x 2 ) + 4 k T f e k x 2 k x T f e k x 2   ,     0 x < 1 ,
where y ( 0 ) > 0 and y ( 1 ) = 0 . Under the condition 0 x < 1 , the derivative of y is obtained as
y ˙ = 8 k e k x T f ( 1 + x 2 ) 2 4 k 2 T f e k x 2 k T f e k x 2 + 4 k 2 x 2 T f e k x 2 ,
y ¨ = 8 k e k T f ( 1 ( 1 + x 2 ) 2 + 4 x 2 ( 1 + x 2 ) 3 ) + 4 k 2 T f [ ( k e k x k x 3 e k x 2 ) + ( 3 x e k x k x 3 e k x 2 ) ] .
We can get that y ¨ > 0 , y ˙ ( 0 ) < 0 , and y ˙ ( 1 ) = 0 when 0 x < 1 . This means that y ˙ is in a trend of monotone increasing, and y ˙ will not be greater than zero because of y ˙ ( 1 ) = 0 . As a result, y is monotonically decreasing and greater than zero in 0 x < 1 . Specifically, it can be concluded that ρ ˙ i < 0 for all 0 t T f . Thus, the proof of Theorem 2 is complete. □
Some studies have been conducted to affirm the property of the proposed PTPPF. The parameters are given as ρ i 0 = 10 and ρ i = 0.01 . In Figure 2, we can see that the variate ρ i converges to the prescribed steady-state error range within a predefined time T f . Figure 3 shows that ρ i ( t ) satisfies ρ ˙ i ( t ) < 0 when 0 t T f and lim t T f ρ i = lim t T f + ρ i = ρ i . The decreasing rate in the transient stage can be increased by selecting a larger value k . The decreasing rate of PTPPF in the transient phase can be adjusted by changing one parameter k on the premise of not changing the convergence time of PPF to reach steady state.
Remark 1. 
The function ρ i has a simple polynomial form. The time to reach ρ i is prescribed explicitly by one parameter T f , which can be adjusted to meet practical application requirements. Unlike the existing performance functions in [30,31,32], the decreasing rate of ρ i can be adjusted by one parameter k when 0 t T f .

3.2. Transformation Method

As is known to all, it is very difficult to design a controller because of the attitude tracking error constraint in Equation (11). To avoid this issue, the following transformation method is adopted to transform the constrained problem into an unconstrained one [20]:
σ e i ( t ) = z 1 i ( t ) ρ i ( t )   ,     i = 1 , 2 , 3 .
In this paper, we define the equivalent unconstrained equation as
e 1 i = tan ( π 2 σ e i ( t ) ) .
Taking the first and second derivatives of e i , there is
e ˙ 1 i = A i ( z 2 i ρ ˙ i ρ i z 1 i ) e ¨ 1 i = P i e ˙ 1 i + A i ( z ˙ 2 i Φ i )       i = 1 , 2 , 3 ,
where A i = π 2 ρ i ( 1 + e 1 i 2 ) , P i = 2 π 2 ρ i ( z 2 i z 1 i ρ ˙ i ρ i ) e 1 i ρ ˙ i ρ i , and Φ i = z 2 i ρ ˙ i ρ i + z 1 i ρ ¨ i ρ i z 1 i ρ ˙ i 2 ρ i 2 . Equation (22) can be rewritten as
e ˙ 1 = A ( z 2 diag ( ρ ˙ i ρ i ) z 1 ) = e 2 e ˙ 2 = P e ˙ 1 + A ( z ˙ 2 Φ )       = P e ˙ 1 + A ( ( R ˙ R 1 + R J 1 Ω J R 1 ) z 2 + R J 1 ( M c + M a i r )         + ( R ˙ R 1 + R J 1 Ω J R 1 ) Λ ˙ C Λ ¨ C Φ ) + D ,
where D = A d , e 1 = [ e 11 ,   e 12 ,   e 13 ] T .
Remark 2. 
After error transformation, the inequality-constrained problem is transformed into an equality unconstrained one. The smooth function 2 π arctan ( e 1 i ) has the properties lim e 1 i + 2 π arctan ( e 1 i ) = 1 and lim e 1 i 2 π arctan ( e 1 i ) = 1 , i = 1 , 2 , 3 . The transformed error function in Equation (21) ensures that the steady-state tracking errors can converge to zero when the transformed error e 1 approaches zero.

4. Nonsingular Predefined-Time Attitude Controller Design

In the traditional predefined-time terminal SMC design process, it can be seen that the negative fractional power term always appears in the control law, which may result in a singularity if e 1 = 0 , e 2 0 [20]. Thus, in this section, a novel nonsingular terminal sliding controller without using piecewise continuous functions is proposed.
Firstly, a novel predefined-time terminal SMS is expressed as follows:
S = sig 2 + 2 a / 3 e 1 + sig 12 + 4 a 3 2 a ( ( 3 + 2 a ) T c 3 π + 2 3 arctan ( 1 / 3 ) e 2 + sig 2 + 2 a / 3 e 1 ) ,
where T c > 0 and 1.5 < a < 0.75 .
Theorem 3. 
The phenomenon of the singularity is avoided. Once the tracking error e 1 and its derivative e 2 make S = [ S 1 ,   S 2 ,   S 3 ] T = 0 , e 1 and e 2 will converge to zero within the predefined time T c .
Proof of Theorem 3. 
When e 1 = [ e 11 ,   e 12 ,   e 13 ] T and e 2 evolve on the expected SMS, we have
sig 1 / 2 a / 3 e 1 i + ( 3 + 2 a ) T c 3 π + 2 3 arctan 1 3 e 2 i + sig 2 + 2 a / 3 e 1 i = 0 ( 3 + 2 a ) T c 3 π + 2 3 arctan 1 3 e 2 i = sig 2 + 2 a / 3 e 1 i + sig 1 / 2 a / 3 e 1 i e 2 i sign ( e 1 i ) = 3 π + 2 3 arctan 1 3 ( 3 + 2 α ) T c ( | e 1 i | 2 + 2 a / 3 + | e 1 i | 1 / 2 a / 3 ) ,
where i = 1 , 2 , 3 . Choosing the Lyapunov function, V i = | e 1 i | and the derivative of V i can be expressed as
V ˙ i = e 2 i sign ( e 1 i )       = 3 π + 2 3 arctan 1 3 ( 3 + 2 α ) T c ( | e 1 i | 2 + 2 a / 3 + | e 1 i | 1 / 2 a / 3 )       3 π + 2 3 arctan 1 3 ( 3 + 2 a ) T c ( V i 1 / 2 a / 3 + V i 2 + 2 a / 3 ) .
By applying Theorem 1, the proposed nonsingular SMS is predefined-time stable within T c . Meanwhile, e 1 and e 2 will reach the SMS S = 0 , whose proof is given in Theorem 4. The proof of Theorem 3 is completed. □
The simulation below affirms the effectiveness of the proposed predefined-time SMS. The different values of a and T c are taken when the initial values satisfy e 1 i = 1 , e 2 i = 0 , and S i = 0 , i = 1 ,   2 ,   3 . In Figure 4, we can see that e 1 i converges to zero within T c .
Then, a predefined-time extended state observer (PTESO) is introduced, which can effectively estimate the disturbances D . The form of PTESO is expressed as follows:
φ 1 = e 1 e ^ 1 e ^ ˙ 1 = e 2 + Ψ ζ Ξ ¯ l 1 φ 1 + ( 1 Ψ ) k 1 ϒ 1 / 3 sig 2 / 3 ( φ 1 ) e ^ ˙ 2 = Δ ^ + Ψ ζ 2 Ξ ¯ 2 l 2 φ 1 + ( 1 Ψ ) k 2 ϒ 2 / 3 sig 1 / 3 ( φ 1 ) + H + U Δ ^ ˙ = Ψ ζ 3 Ξ ¯ 3 l 3 φ 1 + ( 1 Ψ ) k 3 ϒ sig 0 ( φ 1 ) ,
where H = P e ˙ 1 + A ( ( R ˙ R 1 + R J 1 Ω J R 1 ) z 2 Λ ¨ C Φ         + ( R ˙ R 1 + R J 1 Ω J R 1 ) Λ ˙ C ) , U = A R J 1 ( M c + M a i r ) , Ξ ¯ = sec 2 ( π t 2 t ¯ f )   ,     t [ 0 ,   t ¯ f ) , and t ¯ f is the convergence time. The time switching function is
Ψ = { 1 ,   t [ 0 ,   t ¯ f ) 0 ,   t [ t ¯ f ,   + ) .
ϒ is a constant to be designed. k 1 , 2 , 3 are the parameters satisfying k 3 > ϒ , k 1 = 3.34 k 3 1 / 3 , and k 2 = 5.3 k 3 2 / 3 . The stability of the PTESO can be referenced in Cui’s paper [33]. According to the previous work, the proposed PTESO has high disturbance estimation accuracy, i.e., lim t > t ¯ f Δ ^ D = 0 .
Considering the predefined-time terminal SMS in Equation(24), the novel nonsingular controller can be designed as
M c = ( A R J 1 ) 1 ( A ( ( R ˙ R 1 + R J 1 Ω J R 1 ) z 2 + ( R ˙ R 1 + R J 1 Ω J R 1 ) Λ ˙ C Λ ¨ C Φ )         + Θ ( 2 + 2 a 3 ) diag ( | e 1 i | 1 + 2 a / 3 ) ( Γ + e 2 )         + π 2 η T p diag ( | S i | 1 η / 2 + 3 η / 2 | S i | 1 + η / 2 ) sign ( S ) + P e ˙ 1 + Δ ^ ) M a i r ,
Γ = 3 2 a 12 + 4 a Θ sig 8 a 6 3 2 a ( e 2 / Θ + sig 2 + 2 3 a e 1 ) + 3 2 a 12 + 4 a Θ sig ( e 2 / Θ + sig 2 + 2 3 a e 1 ) ,
where Θ = 3 π + 2 3 arctan 1 3 ( 3 + 2 a ) T c , i = 1 , 2 , 3 . 0 < η < 1 , T p > t ¯ f > 0 .
The controller in Equation (29) and the auxiliary function Γ are nonsingular because 1.5 < a < 0.75 . By combining Equation (24) and Equation (29), the following equation is established:
12 + 4 a 3 2 a | e 2 / Θ + sig 2 + 2 3 a e 1 | 9 + 6 a 3 2 a Γ = e 2 + Θ S .
Theorem 4. 
Under Assumption 1, if the controller is designed as Equation (29), then e 1 and e 2 will reach S = 0 .
Proof of Theorem 4. 
Firstly, differentiating S leads to
S ˙ = ( 2 + 2 3 a ) diag ( | e 1 i | 1 + 2 a / 3 ) e 2 + Φ ( e ˙ 2 Θ + ( 2 + 2 a 3 ) diag ( | e 1 i | 1 + 2 a / 3 ) e 2 )       = ( 2 + 2 3 a ) diag ( | e 1 i | 1 + 2 a / 3 ) e 2 + Φ ( 1 Θ ( P e ˙ 1 + A ( ( R ˙ R 1 + R J 1 Ω J R 1 ) z 2 + R J 1 ( M c + M a i r )         + ( R ˙ R 1 + R J 1 Ω J R 1 ) Λ ˙ C Λ ¨ C Φ ) + D ) + ( 2 + 2 a 3 ) diag ( | e 1 i | 1 + 2 a / 3 ) e 2 )       = ( 2 + 2 3 a ) diag ( | e 1 i | 1 + 2 a / 3 ) e 2 + Φ ( 1 Θ ( Θ ( 2 + 2 a 3 ) diag ( | e 1 i | 1 + 2 a / 3 ) ( Γ + e 2 )         + π 2 η T p diag ( | S i | 1 η / 2 + 3 η / 2 | S i | 1 + η / 2 ) sign ( S ) Δ ^ + D )     + ( 2 + 2 a 3 ) diag ( | e 1 i | 1 + 2 a / 3 ) e 2 )       = ( 2 + 2 3 a ) diag ( | e 1 i | 1 + 2 a / 3 ) ( Θ sig 2 + 2 a / 3 ( e 1 ) + Θ sig 12 + 4 a 3 2 a ( 1 Θ e 2 + sig 2 + 2 a / 3 ( e 1 ) ) )             + 1 Θ Φ ( π 2 η T p diag ( | S i | 1 η / 2 + 3 η / 2 | S i | 1 + η / 2 ) sign ( S ) Δ ^ + D )       = 1 Θ Φ ( π 2 η T p diag ( | S i | 1 η / 2 + 3 η / 2 | S i | 1 + η / 2 ) sign ( S ) Δ ^ + D ) ( 2 + 2 3 a ) diag ( | e 1 i | 1 + 2 a / 3 ) S ,
where Φ = 12 + 4 a 3 2 a diag ( | e 2 i Θ + sig 2 + 2 a / 3 ( e 1 i ) | 9 + 2 a 3 2 a ) , i = 1 ,   2 ,   3 .
Choosing the Lyapunov function yields
V S = | S 1 | + | S 2 | + | S 3 | ,
where S = [ S 1 ,   S 2 ,   S 3 ] T . By differentiating Equation (33), on the basis of Lemma 2, we have
V ˙ S = | S 1 | + | S 2 | + | S 3 |       = sign ( S ) T S ˙       = sign ( S ) T ( 1 Θ Φ ( π 2 η T p diag ( | S i | 1 η / 2 + 3 η / 2 | S i | 1 + η / 2 ) sign ( S ) Δ ^ + D )             ( 2 + 2 3 a ) diag ( | e 1 i | 1 + 2 a / 3 ) S )       min ( 1 Θ Φ i i ) π 2 η T p ( V S 1 η / 2 + V S 1 + η / 2 ) + 1 Θ sign ( S ) T Φ ( Δ ^ + D ) .
According to Equation (27), the disturbances D can be precisely estimated by PTESO when t ¯ f < t < T p . Specifically, the following inequality holds.
V ˙ S min ( 1 Θ Φ i i ) π 2 η T p ( V S 1 η / 2 + V S 1 + η / 2 )         1 Θ Φ j j π 2 η T p ( V S 1 η / 2 + V S 1 + η / 2 )         G ( π 2 η T p ( V S 1 η / 2 + V S 1 + η / 2 ) ) ,
where i = 1 , 2 , 3 . j = i when Φ j j = min ( Φ i i ) . Then, we define ψ = e 2 j Θ + sig 2 + 2 a / 3 ( e 1 j ) , G = 1 Θ Φ j j = 12 + 4 a 3 2 a ( 3 + 2 a ) T c 3 π + 2 3 arctan 1 3 | ψ | 9 + 2 a 3 2 a . Thus, the state spaces are divided into two areas, i.e., Ω 1 = { ( e 1 j ,   e 2 j ) | G 1 } and Ω 2 = { ( e 1 j ,   e 2 j ) | G < 1 } . If the system states are in Ω 1 , we have
V ˙ S G ( π 2 η T p ( V S 1 η / 2 + V S 1 + η / 2 ) )         π 2 η T p ( V S 1 η / 2 + V S 1 + η / 2 ) .
According to Lemma 1, the attitude tracking errors will reach S = 0 or Ω 2 within a predefined time. If the system states are in Ω 2 when ψ 0 , the SMS S = 0 is still an attractor. The remainder of the study demonstrates that the curve ψ = 0 is unattractive except for S = 0 . There is Γ = 0 when ψ = 0 , and the control command M c is described as
M c = ( A R J 1 ) 1 ( A ( ( R ˙ R 1 + R J 1 Ω J R 1 ) z 2 + ( R ˙ R 1 + R J 1 Ω J R 1 ) Λ ˙ C Λ ¨ C Φ )         + Θ ( 2 + 2 a 3 ) diag ( | e 1 i | 1 + 2 a / 3 ) e 2         + π 2 η T p diag ( | S i | 1 η / 2 + 3 η / 2 | S i | 1 + η / 2 ) sign ( S ) + P e ˙ 1 + Δ ^ ) M a i r .
Differentiating ψ yields
ψ ˙ = π 2 η T p ( | S j | 1 η / 2 + 3 η / 2 | S j | 1 + η / 2 ) sign ( S j ) .
It is obvious that ψ ˙ < 0 for S j > 0 and ψ ˙ > 0 for S j < 0 . Therefore, the curve ψ = 0 is not an attractor except for S = 0 ; the attitude tracking errors will transgress the area Ω 2 into Ω 1 monotonically. □
Remark 3. 
The transient and steady-state performance of the attitude tracking errors can be sufficiently guaranteed under the proposed PTPPF. Compared with the traditional nonsingular predefined-time controller [34,35], the singularity can be avoided without using piecewise continuous functions.
Remark 4. 
It is theoretically possible to design a very short convergence time for RLV attitude control, but this will lead to a very large control moment. Therefore, in practical application, the convergence time needs to be reasonably designed according to the mission requirement and the ability of RLV control system.

5. Simulation and Analysis

The simulation of a 6-DOF RLV dynamic is conducted in this section. The parameters of RLV are J x x = 130700   kg m 2 , J y y = 1507000   kg m 2 , J z z = 1493000   kg m 2 , and J x z = 116000   kg m 2 . The cross-section area and reference length are S ref = 73   m 2 and L r = 40   m , respectively. The initial state variables in the simulation are given as follows: r = 6457   km , θ = 17.32 , φ = 105.01 , σ = 0 . 01 , β = 0.49 , α = 45.01 , p = 0 . 0007   rad / s , q = 0.0   rad / s , r = 0.0698   rad / s , v = 2916   m / s , and m = 23000   kg . Moreover, the parameters of PTPPF are set as ρ i 0 = 0.2 rad , ρ i = 0.005 rad , k = 1.6 , and T f = 1 . The parameters of the controller are a = 1.2 , T c = 0.5 , n = 0.5 , and T p = 0.5 . The parameter uncertainties of aerodynamic coefficients in the simulation are set to 20% bias, while 10% bias is set for the moment of inertia. Then, two different control simulations are conducted for comparison; one is for the fixed-time nonsingular fast terminal sliding mode control (FxNFTSMC) [14], and the other is for active disturbance rejection controller (ADRC) [36]. For FxNFTSMC, let v 1 = 1.1 , v 2 = 3.2 , λ 1 = 1.6 , λ 2 = 1.3 , γ 1 = 1.3 , γ 2 = 0.38 , χ = 1.5 , and λ = 2.0 . In the ADRC, there is β 1 = ( 150 ,   160 ,   160 ) , β 2 = ( 9.09   ×   10 3 ,   9.09   ×   10 3 ,   9.09   ×   10 3 ) , β 3 = ( 1   ×   10 5 ,   2   ×   10 3 ,   2   ×   10 3 ) , c = ( 0.4 ,   0.4 ,   0.4 ) , r = ( 810 ,   830 ,   305 ) , and h 0 = ( 0.1 ,   0.1 ,   0.1 ) .
In Figure 5, Figure 6, Figure 7 and Figure 8, we can see that the attitudes of RLV can track the time-varying guidance command Λ C in the desired time 1.0 s, and the deviation from guidance commands is less than the predefined value ρ i = 0.005   rad . Figure 5 shows that the proposed controller has an excellent control effect taking into account the parameter uncertainties, state constraints, and external disturbances. In contrast, the attitude tracking errors of the FxNFTSMC are eliminated to be near zero within 1.45 s, and the convergence time of ADRC is about 1.5 s. Meanwhile, the FxNFTSMC needs a complex calculation of design parameters to obtain the desired convergence time. The proposed controller can directly give the actual convergence time, which avoids the process of obtaining the convergence time of the system through constant iteration and repeated testing in the FxSMC.
Figure 7 shows the curve of the SMS, and the control moment is shown in Figure 8. It can be seen that the proposed method can effectively alleviate the chattering phenomenon. The control moment becomes larger at 25 s of simulation time, which is due to a sudden change in the rate of the angle of attack.
In Figure 9 and Figure 10, the value of ρ continuously changes to ρ in the prescribed time T f . This indicates the effectiveness of the controller, which can prevent z 1 from exceeding the prescribed bounds.

6. Conclusions

A novel predefined-time prescribed performance nonsingular sliding mode controller was designed for attitude control of VTHL RLV. A PTPPF is put forward to guarantee that the attitude tracking errors are constrained within the desired transient and steady-state boundaries. On the basis of the error transformation method, the novel predefined-time nonsingular controller can guarantee that the attitude tracking errors converge to the predefined steady-state boundary in a predefined time. Meanwhile, the problem of singularity is avoided without using piecewise continuous functions. The stability of the system was verified using the Lyapunov method. The performance of the proposed method was confirmed using numerical simulations.

Author Contributions

Conceptualization, M.W., C.W. and J.P.; methodology, M.W.; software, M.W.; validation, C.W. and J.P.; resources, N.C.; data curation, J.P.; writing—original draft preparation, M.W.; writing—review and editing, C.W.; visualization, C.W.; supervision, N.C.; project administration, J.P. All authors read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Defense Basic Scientific Research (JCKY2019208C017) and National Natural Science Foundation of China (Grant 52232014).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors would like to thank the editors and anonymous reviewers for their constructive comments and suggestions.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Consider the following equality:
V ˙ = 3 π + 2 3 arctan 1 3 ( 3 + 2 α ) T c ( V 1 / 2 α / 3 + V 2 + 2 α / 3 ) .
Integrating Equation (A1) yields
T ( x 0 ) = V ( x 0 ) 0 d V 3 π + 2 3 arctan 1 3 ( 3 + 2 α ) T c ( V 1 / 2 α / 3 + V 2 + 2 α / 3 )                   = ( 3 + 2 α ) T c 3 π + 2 3 arctan 1 3 V ( x 0 ) 0 d V ( V 1 / 2 α / 3 + V 2 + 2 α / 3 )                   = 6 T c 3 π + 2 3 arctan 1 3 V ( x 0 ) 0 d V 1 / 2 + α / 3 1 + V 3 / 2 + α                   = 6 T c 3 π + 2 3 arctan 1 3 [ 1 3 ln | V 1 / 2 + α / 3 + 1 ( V 1 / 2 + α / 3 ) 2 V 1 / 2 + α / 3 + 1 | + 1 3 arctan ( 2 V 1 / 2 + α / 3 1 3 ) ] | V ( x 0 ) 0 .
It is apparent that the setting time function T ( x 0 ) is less than T c because of lim V ( x 0 ) T ( x 0 ) = T c . Finally, we can conclude that the system in Equation (7) is globally predefined-time stable within predefined-time T c .

References

  1. Dong, Q.; Zong, Q.; Tian, B.; Wang, F. Integrated finite-time disturbance observer and controller design for reusable launch vehicle in reentry phase. J. Aerosp. Eng. 2017, 30, 04016076. [Google Scholar] [CrossRef]
  2. Ghasemi, A.; Moradi, M.; Menhaj, M.B. Adaptive fuzzy sliding mode control design for a low-lift reentry vehicle. J. Aerosp. Eng. 2012, 25, 210–216. [Google Scholar] [CrossRef]
  3. Song, J.; Wang, L.; Cai, G.; Qi, X. Nonlinear fractional order proportion-integral-derivative active disturbance rejection control method design for hypersonic vehicle attitude control. Acta Astronaut. 2015, 111, 160–169. [Google Scholar] [CrossRef]
  4. Ming, C.; Sun, R.; Wang, X.M. Velocity control based on active disturbance rejection for air-breathing supersonic vehicles. Complexity 2018, 3, 6217657. [Google Scholar]
  5. Mao, Q.; Dou, L.Q.; Zong, Q.; Ding, Z. Attitude controller design for reusable launch vehicles during reentry phase via compound adaptive fuzzy H-infinity control. Aerosp. Sci. Technol. 2018, 72, 36–48. [Google Scholar] [CrossRef] [Green Version]
  6. Zhang, X.Y.; Zong, Q.; Dou, L.; Tian, B.; Liu, W. Improved finite-time command filtered backstepping fault-tolerant control for flexible hypersonic vehicle. J. Frankl. Inst. 2020, 357, 8543–8565. [Google Scholar] [CrossRef]
  7. Xu, B.; Wang, X.; Shi, Z.K. Robust Adaptive neural control of nonminimum phase hypersonic vehicle model. IEEE Trans. Syst. Man Cybern. Syst. 2019, 52, 1107–1115. [Google Scholar]
  8. Xu, X.Z.; Wu, W.X.; Zhang, W.G. Sliding mode control for a nonlinear aeroelastic system through backstepping. J. Aerosp. Eng. 2018, 31, 04017080. [Google Scholar] [CrossRef]
  9. Jiang, S.; Liu, C.S.; Gao, Y.X. MIMO adaptive high-order sliding mode control for quadrotor attitude tracking. J. Aerosp Eng. 2021, 34, 04021022. [Google Scholar] [CrossRef]
  10. Lu, K.; Xia, Y. Adaptive attitude tracking control for rigid spacecraft with finite-time convergence. Automatica. 2013, 49, 3591–3599. [Google Scholar] [CrossRef]
  11. Hall, C.E.; Shtessel, Y. Sliding mode disturbance observer-based control for a reusable launch vehicle. J. Guid. Control. Dyn. 2006, 29, 1315–1328. [Google Scholar] [CrossRef]
  12. Wei, C.Z.; Wang, M.Z.; Lu, B.; Pu, J. Accelerated Landweber iteration based control allocation for fault tolerant control of reusable launch vehicle. Chin. J. Aeronaut. 2021, 35, 175–184. [Google Scholar] [CrossRef]
  13. Chen, Y.; Zhu, B. Distributed fixed-time control of high-order multi-agent systems with non-holonomic constraints. J. Franklin Inst. 2021, 358, 2948–2963. [Google Scholar] [CrossRef]
  14. Zhang, L.; Wei, C.; Wu, R.; Cui, N. Fixed-time extended state observer based non-singular fast terminal sliding mode control for a VTVL reusable launch vehicle. Aerosp. Sci. Technol. 2018, 82, 70–79. [Google Scholar] [CrossRef]
  15. Ju, X.Z.; Wang, F.; Wei, C.; Zhang, L. Asymmetric barrier Lyapunov function based practical fixed-time control of vertical take-off/ vertical landing reusable launch vehicles with partial state constraints. IFAC-PapersOnLine 2020, 53, 14845–14850. [Google Scholar] [CrossRef]
  16. Liang, X.H.; Wang, Q.; Hu, C.; Dong, C. Fixed-time observer based fault tolerant attitude control for reusable launch vehicle with actuator faults. Aerosp. Sci. Technol. 2020, 107, 106314. [Google Scholar] [CrossRef]
  17. Sánchez-Torres, J.D.; Sanchez, E.N.; Loukianov, A.G. Predefined-time stability of dynamical systems with sliding modes, in: American Control Conference. In Proceedings of the 2015 American Control Conference, Chicago, IL, USA, 1–3 July 2015; pp. 5842–5846. [Google Scholar]
  18. Lyu, W.; Zhai, D.-H.; Xiong, Y.; Xia, Y. Predefined performance adaptive control of robotic manipulators with dynamic uncertainties and input saturation constraints. J. Frankl. Inst. 2021, 2, 7142–7169. [Google Scholar] [CrossRef]
  19. Sánchez-Torres, J.D.; Muñoz-Vázquez, A.J.; Defoort, M.; Jiménez-Rodríguez, E.; Loukianov, A.G. A class of predefined-time controllers for uncertain second-order systems. Eur. J. Control. 2019, 53, 52–58. [Google Scholar]
  20. Zhang, L.; Ju, X.Z.; Cui, N.G. Ascent control of heavy-lift launch vehicle with guaranteed predefined performance. Aerosp. Sci. Technol. 2021, 110, 106511. [Google Scholar] [CrossRef]
  21. Wu, C.; Yan, J.; Wu, X.; Guo, Y.; Mou, P.; Xiao, B. Predefined-time attitude stabilization of receiver aircraft in aerial refueling. IEEE Trans. Circuits Syst. II Express Briefs 2021, 68, 3321–3325. [Google Scholar] [CrossRef]
  22. Bu, X.; Wu, X.; Zhu, F.; Huang, J.; Ma, Z.; Zhang, R. Novel prescribed performance neural control of a flexible air-breathing hypersonic vehicle with unknown initial errors. ISA Trans. 2015, 59, 149–159. [Google Scholar] [CrossRef]
  23. Zhang, C.; Ma, G.; Sun, Y.; Li, C. Prescribed performance adaptive attitude tracking control for flexible spacecraft with active vibration suppression. Nonlinear Dyn. 2019, 96, 1909–1926. [Google Scholar] [CrossRef]
  24. Luo, J.; Yin, Z.; Wei, C.; Yuan, J. Low-complexity prescribed performance control for spacecraft attitude stabilization and tracking. Aerosp. Sci. Technol. 2018, 74, 173–183. [Google Scholar] [CrossRef]
  25. Tian, B.L.; Fan, W.; Zong, Q.; Wang, J.; Wang, F. Nonlinear robust control for reusable launch vehicles in reentry phase based on time-varying high order sliding mode. J. Frankl. Inst. 2013, 350, 1787–1807. [Google Scholar] [CrossRef]
  26. Munoz-Vazquez, A.J.; Sanchez-Torres, J.D.; Jimenez-Rodriguez, E.; Loukianov, A.G. Predefined-time robust stabilization of robotic manipulators. IEEE ASME Trans. Mechatron. 2019, 24, 1033–1040. [Google Scholar] [CrossRef]
  27. Liu, L.; Yao, W.; Guo, Y. Prescribed performance tracking control of a free-flying flexible-joint space robot with disturbances under input saturation. J. Frankl. Inst. 2021, 358, 4571–4601. [Google Scholar] [CrossRef]
  28. Ye, D.; Zou, A.M.; Sun, Z.W. Predefined-time predefined-bounded attitude tracking control for rigid spacecraft. IEEE T. Aero. Elec. Sys. 2021, 58, 464–472. [Google Scholar] [CrossRef]
  29. Qian, C.J.; Wei, L. A continuous feedback approach to global strong stabilization of nonlinear systems. IEEE Trans. Autom. Control. 2001, 46, 1061–1079. [Google Scholar] [CrossRef]
  30. Bechlioulis, C.P.; Rovithakis, G.A. Robust adaptive control of feedback linearizable MIMO nonlinear systems with prescribed performance. IEEE Trans. Autom. Control 2008, 53, 9. [Google Scholar] [CrossRef]
  31. Zhang, J.; Yang, G. Prescribed performance fault-tolerant control of uncertain nonlinear systems with unknown control directions. IEEE Trans. Autom. Control 2017, 62, 12. [Google Scholar] [CrossRef]
  32. Hu, Q.; Shao, X.; Guo, L. Adaptive fault-tolerant attitude tracking control of spacecraft with prescribed performance. IEEE/ASME Trans. Mechatron. 2018, 23, 1. [Google Scholar] [CrossRef]
  33. Cui, L.; Jin, N. Prescribed-time ESO-based prescribed-time control and its application to partial IGC design. Nonlinear Dyn. 2021, 106, 491–508. [Google Scholar] [CrossRef]
  34. Shao, X.Y.; Sun, G.H.; Xue, C. Nonsingular terminal sliding mode control for free-floating space manipulator with disturbance. Acta Astronaut. 2021, 181, 396–404. [Google Scholar] [CrossRef]
  35. Wang, F.; Miao, Y.; Li, C.Y. Attitude control of rigid spacecraft with predefined-time stability. J. Franklin Inst. 2020, 357, 7. [Google Scholar]
  36. Hu, H.M.; Du, J.W. Active disturbance rejection control for permanent magnet linear synchronous motor. In Proceedings of the 13th International Symposium on Linear Drives for Industry Applications, Wuhan, China, 1–3 July 2021. [Google Scholar]
Figure 1. VTHL reusable launch vehicle.
Figure 1. VTHL reusable launch vehicle.
Applsci 12 10153 g001
Figure 2. The simulation results with different T f and k .
Figure 2. The simulation results with different T f and k .
Applsci 12 10153 g002
Figure 3. The simulation results of PTPPF: (a) the curves of ρ ˙ i for five cases; (b) the curves of ρ ¨ i for five cases.
Figure 3. The simulation results of PTPPF: (a) the curves of ρ ˙ i for five cases; (b) the curves of ρ ¨ i for five cases.
Applsci 12 10153 g003
Figure 4. The simulation results of the tracking errors.
Figure 4. The simulation results of the tracking errors.
Applsci 12 10153 g004
Figure 5. Attitude tracking curves of σ ,   β ,   α : (a) the roll channel; (b) the yaw channel; (c) the pitch channel.
Figure 5. Attitude tracking curves of σ ,   β ,   α : (a) the roll channel; (b) the yaw channel; (c) the pitch channel.
Applsci 12 10153 g005
Figure 6. Time responses of angular velocity: (a) the roll channel; (b) the yaw channel; (c) the pitch channel.
Figure 6. Time responses of angular velocity: (a) the roll channel; (b) the yaw channel; (c) the pitch channel.
Applsci 12 10153 g006
Figure 7. Sliding mode surface of three channels: (a) the roll channel; (b) the yaw channel; (c) the pitch channel.
Figure 7. Sliding mode surface of three channels: (a) the roll channel; (b) the yaw channel; (c) the pitch channel.
Applsci 12 10153 g007
Figure 8. Control moment of three channels: (a) the roll channel; (b) the yaw channel; (c) the pitch channel.
Figure 8. Control moment of three channels: (a) the roll channel; (b) the yaw channel; (c) the pitch channel.
Applsci 12 10153 g008
Figure 9. Attitude tracking errors of σ , β , α .
Figure 9. Attitude tracking errors of σ , β , α .
Applsci 12 10153 g009
Figure 10. Attitude velocity errors.
Figure 10. Attitude velocity errors.
Applsci 12 10153 g010
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, M.; Wei, C.; Pu, J.; Cui, N. Predefined-Time Nonsingular Attitude Control for Vertical-Takeoff Horizontal-Landing Reusable Launch Vehicle. Appl. Sci. 2022, 12, 10153. https://doi.org/10.3390/app121910153

AMA Style

Wang M, Wei C, Pu J, Cui N. Predefined-Time Nonsingular Attitude Control for Vertical-Takeoff Horizontal-Landing Reusable Launch Vehicle. Applied Sciences. 2022; 12(19):10153. https://doi.org/10.3390/app121910153

Chicago/Turabian Style

Wang, Mingze, Changzhu Wei, Jialun Pu, and Naigang Cui. 2022. "Predefined-Time Nonsingular Attitude Control for Vertical-Takeoff Horizontal-Landing Reusable Launch Vehicle" Applied Sciences 12, no. 19: 10153. https://doi.org/10.3390/app121910153

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